What Is a Complex System?
What Is a Complex System?
J. Ladyman and K. Wiesner
New Haven & London
Copyright © 2020 by James Ladyman and Karoline Wiesner.
All rights reserved.
This book may not be reproduced, in whole or in part, including
illustrations, in any form (beyond that copying permitted by Sections 107
and 108 of the U.S. Copyright Law and except by reviewers for the public
press), without written permission from the publishers.
Yale University Press books may be purchased in quantity for educational,
business, or promotional use. For information, please e-mail
sales.press@yale.edu (U.S. office) or sales@yaleup.co.uk (U.K. office).
Printed in the United States of America.
Library of Congress Control Number: 2020930179 ISBN 978-0-300-25110-
4 (pbk: alk. paper)
A catalogue record for this book is available from the British Library.
This paper meets the requirements of ANSI/NISO Z39.48-1992
(Permanence of Paper).
10 9 8 7 6 5 4 3 2 1
Contents
Preface
1 Introduction
1.1 What Is a Complex System?
1.2 A Brief History of Complexity Science
1.2.1 Cybernetics and Systems Theory
1.2.2 Dynamical Systems Theory
1.2.3 Cellular Automata
1.2.4 The Rise of Complexity Science
2 Examples of Complex Systems
2.1 Matter and Radiation
2.2 The Universe
2.3 The Climate System
2.4 Eusocial Insects
2.4.1 Ant Colonies
2.4.2 Honeybee Hives
2.5 Markets and Economies
2.6 The World Wide Web
2.7 The Human Brain
3 Features of Complex Systems
3.1 Numerosity
3.2 Disorder and Diversity
3.3 Feedback
3.4 Non-Equilibrium
3.5 Interlude: Emergence
3.6 Order and Self-Organisation
3.7 Nonlinearity
3.8 Robustness
3.9 Nested Structure and Modularity
3.10 History and Memory
3.11 Adaptive Behaviour
3.12 Different Views of Complexity
4 Measuring Features of Complex Systems
4.1 Numerosity
4.2 Disorder and Diversity
4.3 Feedback
4.4 Non-Equilibrium
4.5 Spontaneous Order and Self-Organisation
4.6 Nonlinearity
4.6.1 Nonlinearity as Power Laws
4.6.2 Nonlinearity versus Chaos
4.6.3 Nonlinearity as Correlations or Feedback
4.7 Robustness
4.7.1 Stability Analysis
4.7.2 Critical Slowing Down and Tipping Points
4.7.3 Self-Organised Criticality and Scale Invariance
4.7.4 Robustness of Complex Networks
4.8 Nested Structure and Modularity
4.9 History and Memory
4.10 Computational Measures
4.10.1 Thermodynamic Depth
4.10.2 Statistical Complexity and True Measure Complexity
4.10.3 Effective Complexity
4.10.4 Logical Depth
5 What Is a Complex System?
5.1 Nihilism about Complex Systems
5.2 Pragmatism about Complex Systems
5.3 Realism about Complex Systems
5.3.1 Generic Conceptions of Complexity
5.3.2 Physical Conceptions of Complexity
5.3.3 Computational Conceptions of Complexity
5.3.4 Functional Conceptions of Complexity
5.4 The Varieties of Complex Systems
5.5 Implications
Appendix – Some Mathematical Background
A Probability Theory
B Shannon Information Theory
C Algorithmic Information Theory
D Complex Networks
Bibliography
Index
Preface
This book is the result of a decade-long collaboration that began in 2007
with the opening of the Centre for Complexity Sciences at the University of
Bristol (an EPSRC-funded doctoral training centre). Teaching at this centre
brought us together, and we began discussing the different phenomena and
measures of complexity science and the ideas associated with it. We
realised the need for a thorough analysis which would answer questions
such as ‘Is complexity a truly new phenomenon or merely a new label?’;
‘Can the different conceptions of complexity of physicists, biologists, social
scientists and others be brought into a single framework, or do they address
different and unrelated phenomena?’; ‘Are measures of complexity
meaningful for a phenomenon this multi-faceted?’; and ‘Why are
information theory and network theory so prominent in complexity
science?’. We strongly believed that, whatever the answers to these
questions are, it would be beneficial to define precisely the terms of the
debate, the phenomena they describe, and the relations between these
phenomena. Lack of clarity in these respects is detrimental to science, and
confused foundations are sometimes highly problematic. We came to the
conclusion that the field of complexity sciences has harboured a lot of
confusion, perhaps because it is rather young and includes so many
different branches. Our first work on this problem led to Ladyman,
Lambert, and Wiesner (2013), upon which this book builds.
This book is written for students and academics interested in complexity
science and the nature of complexity and for other scientific practitioners in
related areas, as well as for scientifically informed general readers. We have
striven for conceptual and linguistic clarity and precision throughout and
have sought to make our ideas and reasoning as simple as possible, while
always being scientifically accurate. We explain both the foundations of
complexity and the mathematical and computational tools currently being
used by complexity scientists.
The first chapter provides an overall introduction to the subject and a
brief account of its history and that of related fields. This chapter is very
widely accessible and contains much that will be familiar to complexity
scientists, though even they will find it worthwhile to read how we
formulate the ideas and issues. We have found that sense can be made of
most of what has been written by complexity scientists about complexity
and that most of what is claimed is for good reason. However, sense and
reason can be obscured by the words used to express them. We hope some
of what we say in this book seems obvious in hindsight because it
formulates clearly and exactly what many people already know. For
example, the ‘truisms’ of complexity science are unlikely to be disputed by
any expert in complexity science, though they have never been stated this
clearly and explicitly before.
The second chapter reviews typical examples of complex systems and
shows how diverse they are and the wide range of features that they display,
as well as some of their commonalities. This chapter is also largely very
widely accessible, though some parts presuppose some knowledge of
physical science. Again we think that experts will find much of what we say
uncontroversial, though some will dispute that all the examples we discuss
are genuine examples of complex systems. Our method in the rest of the
book is to answer the questions raised briefly above and more fully in the
first chapter by using the examples we discuss in the second chapter as data
for our conceptual analysis.
Chapter 3 and Chapter 4 are the core of our account of the foundations of
complexity. In the former we compose a list of features of complex
systems, and in the latter we relate this to the measures used by scientists in
the field. Some of the material in Chapter 4 requires knowledge of
mathematics as we explain many mathematical and computational tools and
their role in studying complexity and refer to the relevant scientific
literature. The final chapter argues for our own view of the consequences of
the analysis of the preceding chapters for the notion of a complex system
and the phenomena of complexity, as well as for the status of complexity
science as a discipline. Parts of this chapter rely on the discussions of the
previous chapters, but we summarise their conclusions so that the reader
who did not follow all the details can still follow our reasoning. We do not
engage much directly with the extensive philosophical literature on
emergence and reduction as that would require a book unto itself.
Philosophers have defined many versions of both ideas, and we have
adopted the most simple taxonomy as explained in the text.
We have many people to thank for their help and support over the years,
for their insightful comments on drafts of this book, and for discussions of
the subject. In particular, we would like to thank the students of the Bristol
Centre for Complexity Sciences for many discussions on the subject and for
their feedback on early drafts of this book. We would like to thank our
colleagues at the University of Bristol and elsewhere. Colleagues who have
been particularly helpful with comments on the manuscript are Colm
Connaughton, Doyne Farmer, Mauro Fazio, Alasdair Houston, Jenann
Ismael, Christopher Jones, Gordon McCabe, Melanie Mitchell, Samir
Okasha, Stuart Presnell, Don Ross, Anne-Lene Sax, Danny Schmidt, Karim
Thébault, Lucy Viegas, Thorsten Wagener, Jim Weatherall, and Lena
Zuchowski as well as a number of anonymous referees. We thank Elisa
Bozzarelli for designing the Figure in Chapter 5.
Chapter 1
Introduction
Complexity science is relatively new but already indispensable. It is
important to understand complex systems because they are everywhere.
Your brain is a complex system and so is your immune system and every
cell in your body. All living systems and all intelligent systems are complex
systems. The climate of the Earth is a complex system, and even the
universe itself exhibits some of the features of complex systems. Many of
the most important problems in engineering, medicine and public policy are
now addressed with the ideas and methods of complexity science – for
example, questions about how epidemics develop and spread. Thousands of
years of mathematical and scientific study have given us the technology to
create new complex systems that rival those of the biosphere, such as cities,
financial economies and the Internet of Things. Business leaders have
started to think in terms of complexity science, using terms such as
‘robustness’, ‘redundancy’ and ‘modularity’ (Sargut and McGrath 2011;
Sullivan 2011). State economic institutions such as the Bank of England
(Haldane 2009) have also begun to use such terminology. This book is
about how scientists think about complex systems and about what makes
these systems special.
However, there is confusion in some of the discussions in the
professional and scientific literature, and clarity is needed to facilitate the
application of complexity science to problems in science and society. There
is no agreement about the definition of ‘complexity’ or ‘complex system’,
nor even about whether a definition is possible or needed. The conceptual
foundations of complexity science are disputed, and there are many and
diverging views among scientists about what complexity and complex
systems are. Even the status of complexity as a discipline can be questioned
given that it potentially covers almost everything.
Most sciences admit of informative definitions that are easy to state. For
example, biology is the study of living systems, chemistry is the study of
molecular structure and its transformations, economics is the study of the
allocation of scarce resources that have different possible uses, and physics
is the study of the most basic behaviour of matter and radiation. Complexity
science is the study of complex systems, and, while it may be difficult
exactly to define ‘life’, ‘matter’ and the other things just mentioned, to say
what complex systems are is even harder. There is no agreement about what
complexity is, whether it can be measured, and, if so, how, and there is no
agreement about whether complex systems all have some common set of
properties.
There are examples that everyone agrees are complex systems, but there
are also many disputed cases. For example, some people regard a purely
physical system like the solar system as a complex system (Simon 1976),
while others think that complex systems must display adaptive behaviour
(Holland 1992; Mitchell 2011), so only systems that have functions and
goals can be complex.1 The rest of this section states clearly what can be
said about complexity science that is not contentious, beginning with the
limitations of the rest of science that make it necessary.
Knowledge of physics and chemistry has enabled us to control many
aspects of the world. The fundamental laws of mechanics and
electromagnetism have a beautiful simplicity and incredible predictive
accuracy. The atomic theory of matter, according to which all the material
things we see around us are composed of elements like carbon and oxygen,
can be used to understand the physical components of every chemical
substance. However, many phenomena are very messy, and the behaviour of
many systems, even relatively simple ones, is very hard to describe in
detail. For example, the flow of turbulent water and the formation of a snow
crystal are incredibly intricate phenomena involving a huge number of
variables (a single snow crystal contains around 1018 molecules). Although
fantastic progress has been made in computation and simulation, measuring
and calculating the state of every molecule in a real snow storm is not
remotely feasible.
Furthermore, the physics and chemistry of atoms and molecules cannot
be used to predict individual people’s actions, where the stock market will
be tomorrow, or what the weather will be next week, because they cannot
be directly applied to such problems at all. People, markets,,the atmosphere,
and their properties are described by psychology, economics and
climatology respectively. Even within physics there are many levels of
description of entities and processes at very different length and time scales,
from the protons and electrons in the standard model of particles, to stars
and galaxies in astrophysics. There is a lot of science that links the
phenomena at different scales. For example, quantum chemistry links
chemical reactions to the electromagnetic interactions between subatomic
particles, and the kinetic theory of gases links the pressure and temperature
of gases to the collisions and motions of their molecules. However, it is
impossible to describe the solar system just using fundamental physics.
In general, collections of things can have different kinds of properties to
their parts. For example, properties like pressure do not pertain to individual
molecules but to gases. In a macroscopic sample of a gas, there are billions
upon billions of molecules and many collisions and motions. If the gas is in
a sealed container, then all these processes automatically make the gas
approximately obey three laws. One of them is Boyle’s law, stating that the
pressure is inversely proportional to the volume at a fixed temperature.
These ‘ideal’ gas laws relate the properties of pressure, volume and
temperature independently of the kind of gas and regardless of the exact
and incredibly complicated behaviour of the particles, all of which are
extremely fast and short-lived compared to the time scale of the behaviour
of the whole gas. (They are called ‘ideal’ because real gases do not obey
them exactly.) Sometimes systems obey laws that are general and allow us
to neglect almost all the details, and in this way simplicity can come from
something very complicated.
There is no need to believe that some mysterious new ingredient has to
be added to molecules to make gases. Gases and their properties are the
result of the relations and interactions among the parts of the gas. If the sum
of the parts is taken to be just the collection of the parts as if they were in
isolation from each other, then the whole is more than the sum of the parts.
However, the interactions of the parts are all it takes to make the whole
exist. One of the most fundamental ideas in complexity science is that the
interactions of large numbers of entities may give rise to qualitatively new
kinds of behaviour different from that displayed by small numbers of them,
as Philip Anderson says in his hugely influential paper, ‘more is different’
(1972).
When whole systems spontaneously display behaviour that their parts do
not, this is called ‘emergence’. Even relatively simple physical systems,
such as isolated samples of gases, liquids and solids, display emergent
phenomena in the minimal sense that they have properties that none of their
individual molecules have singly or in small numbers. However, there are
many different kinds of emergence that are much more intricate – for
example, when systems undergo ‘phase transitions’, such as turning from
liquid to solid or from insulator to superconductor. Phase transitions and
associated ‘critical phenomena’ are examples of spontaneous self-
organisation, in which physical systems are driven from the outside and
there is emergent order to their behaviour. Systems can be driven by heat,
for example, and also by a flow of matter. The famous Belousov-
Zhabotinsky reaction produces patterns of different coloured chemicals that
oscillate as long as more reagents are added. Such examples show that there
are many rich forms of emergent behaviour in nonliving systems and that
nonliving systems can generate order.2 (These examples and those of some
living systems mentioned below are explained in Chapter 2.)
Biological systems display many further examples of emergence,
including metabolism and the coding for proteins in DNA, the
representation of the state of the environment by perceptual systems, and
adaptive behaviour like foraging and the rearing of offspring. Emergence in
collectives of organisms includes the social behaviour found, for example,
in beehives and ant colonies, which are in some ways like a single meta-
organism, and elephant herds and primate groups, whose societies can be
very sophisticated. There are cases of collective motion being directed by a
privileged individual, such as a herd of horses following the leading mare.
However, a flock of birds moves as a whole without a special individual
leading it. Similarly, when social insects make decisions, such as bees
collectively flying off to a new nest, they do so without one individual
playing any special role in the group. Instead, their collective behaviour
arises just as a result of their interactions and the feedback between their
responses to each other’s behaviour. A central idea in complexity science is
that complex systems are spontaneous products of their parts and the
interactions among them. Individual ants and small numbers of them just
wander around aimlessly, but in large numbers they build bridges, maintain
their nests and even grow fungi in them. This is another of the lessons of
complexity science. Coordinated behaviour does not require an overall
controller.
There is sometimes an underlying simplicity to the production of
coordination and order that can be put in mathematical terms. It is
surprising that the collective motion of a flock of birds, a shoal of fish, or a
swarm of insects can be produced by a collection of robots programmed to
obey just a couple of simple rules (Hamann 2018). Each individual must
stay close to a handful of neighbours and must not bump into another
individual. It regularly checks how close it is to others as it moves and
adjusts its trajectory accordingly. As a result, a group moving together
spontaneously forms. The adaptive behaviour of the collective arises from
the repeated interactions, each of which on its own is relatively simple. This
is another conclusion of complexity science: complexity can come from
simplicity.3
The ideal gas laws mentioned above as a very simple example of
emergent behaviour are of limited application because they do not apply to
real gases under many circumstances (for example, at very low
temperatures or very high density or if a gas is compressed very quickly,
heated very rapidly, or suddenly allowed to expand). Similarly, all sciences
involve ways of approximating, idealising and neglecting details. For
example, the law of the pendulum, which says that the time period of
oscillation depends on the length of the string but not on the mass of the
bob, applies only when the line connecting the bob to the pivot can be
treated as being massless because it is so small compared to the mass of the
bob. Similarly, Newton was able to work out the inverse square law of
gravitation only because there is negligible friction to affect the motion of
the planets, and their attraction for each other is negligible compared to the
attraction of the sun. Even in such ideal circumstances, the equations
describing how more than two bodies behave in general cannot be solved
exactly and numerical methods must be used, or a restricted class of
systems must be studied (Goldstein 1950).
Knowing how to model any complex system requires knowing what
idealisations and approximations to make. Complexity science involves
distinctive kinds of approximation and idealisation. For example, the
Schelling model of segregation treats a population and its residences as a
lattice of squares, each of which can be populated or not by one of two
types of individuals (Schelling 1969). The system evolves according to the
rule that individuals move on a given turn if and only if they are surrounded
by fewer individuals of the same type than some specified number. The
stable states of such systems are highly segregated, and in them most
individuals are surrounded by others of the same type. These models show
that segregation can arise even when individuals have a relatively mild
preference for being near others they perceive to be in some way similar to
themselves. This model can be applied not just to residence, but also to the
formation of social networks (Henry et al. 2011).
Sometimes multiple approximations can be made, and different models
of the same system often suit different purposes. For example, the nucleus
of an atom can be modelled as a liquid drop, for the purpose of studying its
overall dynamics, or with its component particles occupying shells
analogous to those used to describe the atomic orbitals of electrons, for
studying how it interacts with radiation. Similarly, there are very diverse
models in complexity science. For example, economic agents can be
modelled as computational agents whose states are updated according to
rules describing flows of information or as nodes in a network that are
connected if two agents trade with each other. Complex systems are often
modelled as networks or information-processing systems.
Complex networks can represent vastly different types of systems and
the connections in a network may represent interactions of various kinds.
For example, both the human body and a city can be modelled as a network
with links representing the flow of energy, food and waste between many
sites. However, networks do not represent only the flow of matter or energy,
but also of information, causal influence, communication, services, or
activation (among other things). In network models, the exact nature of the
interactions may even be ignored when the properties of the system that are
directly studied are the connections and interactions among the parts
considered abstractly (Easley and Kleinberg 2010). In the biological and
behavioural sciences models can be highly abstract – for example, graphs
that only show ancestry relations – and highly idealised – for example,
models of markets that treat agents as having perfect information.
While the interactions between the components in a network have some
particular nature and are governed by the corresponding laws, often we can
ignore the details about them, because the complex behaviour depends only
on more abstract features of the interactions, such as how often they happen
and between which parts. For example, in an economy, agents interact
either face to face, or by post, or electronically, but how they interact is
irrelevant beyond the implications for the timing and reliability of the
exchange of information and resources. Similarly, each bird in a flock is an
individual organism with a heart, a skin, eyes; it has an age, a certain size,
and the need for food for survival and for procreation and many other
things. But when scientists are studying collective motion, all that needs to
be modelled is that the individuals in the group have a way of telling how
close they are to each other. It is not important whether they do so by sight,
like birds, or echolocation, like bats. The effect is the same, as long as they
get the information somehow. Bees communicate by dancing when
choosing where to make a new nest, but that is not important to the model
of how the decision making occurs. Amazingly, the way your brain makes
simple decisions is very similar, with neurons being analogous to bees.
Such similarity is often captured by a common mathematical description of
the different systems in question (more of this in Chapter 2 and Chapter 4).
This is another important lesson of complexity science: There are various
kinds of universality and forms of universal behaviour in complex systems.
Some complex systems involve billions upon billions of interactions
between vast numbers of individuals. The complexity that can emerge is
astonishing. Even the dynamics of the interactions of a thousand birds in a
flock following two simple rules are beyond what a human being can
calculate. Successful scientific modelling of the structure that can arise
from repeated interaction requires computers. Without very powerful
computers, it is impossible, for example, to collate all the data to map the
flow of gas, electricity, water, people, and information in a city. Only for a
few decades have we had the necessary computational power to analyse
complex behaviour, simulate complex systems, and test hypotheses about
how simple interaction rules and feedback produce complex behaviour.
Even with vast computational power many complex systems are so
complicated that making precise predictions about exactly what a particular
system will do is practically impossible. Hence, predictions of real world
complex systems are always of a statistical nature. In general, complexity
science is computational and probabilistic.
Complexity science is often contrasted with reductive science, where the
latter is based on breaking wholes into parts. This is misleading, because, as
the rest of this book shows, complexity science always involves describing
a system by describing the interactions and relations among its parts. The
parts of complex systems interact by various mechanisms studied by
individual scientific disciplines. Furthermore, the remarkable properties of
complex systems arise because of the effects of the laws that govern the
parts and their interactions. However, when there are many parts and they
interact a lot, studying them requires other methods as well as those of the
more fundamental science or sciences that describe the parts and involves
new concepts and theories to describe the novel properties that the parts on
their own do not display. In most complex systems, the interactions between
the parts are of more than one kind. For example, there are both chemical
and electrical interactions in the brain and both electromagnetic and
gravitational interactions in galaxies. Hence, for these reasons, in
complexity science often no single theory encompasses the system of
interest.
Clearly complexity science would not be possible without the rest of
science, and it involves combining theories from different domains and
synthesising tools from various sciences. Complexity science does not
involve revisions to fundamental laws, but it does involve the discovery of
completely new implications of these laws for the behaviour of aggregates
of systems that obey them. This is one reason why complexity science
involves multiple disciplines. Scientific theories that have been studied and
applied autonomously are integrated in a single context. Complexity
science is therefore essentially interdisciplinary in both method and subject
matter. It uses established scientific theories from whatever domain is
relevant to the system at hand and then uses whatever resources are needed
to combine them. Particular sciences provide different aspects of the
explanation of the overall behaviour of the system. The relevant theories
and the relations between them provide the basis for a new (complexity)
theory of the system and new ways of explaining and predicting its features.
Complexity science combines the science specific to the kind of system
being studied with mathematical theories, models and techniques from
computer science, dynamical systems theory, information theory, network
analysis and statistical physics. Much has been learned in this way about
complex systems in neuroscience, cell biology, economics, astrophysics and
many other sciences, and the techniques of complexity science are now
essential for much of engineering, medicine and technology.
Understanding the nature of complex systems is made more difficult by
the fact that complexity science studies both systems that produce complex
structures and those structures themselves. Nature is full of beautiful
patterns and symmetries, such as those of honeycombs, shells and
spiderwebs, which are made by living systems. Intricate structures are also
found in the nonliving world – for example, in the rings of Saturn or
geometrical rock formations on Earth. There is a difference between the
order that complex systems produce and the order of the complex systems
themselves.
The most astonishing example of novel properties arising in a biological
system is the human brain. Our mental life and consciousness somehow
emerge from the electrical and biochemical interactions among neurons.
Human beings and culture are the most complex systems of which we
know, and there are layers upon layers of complexity within them: for
example, the many individual actions that give rise to the single event of an
election or a stock market crash; the intricate feedback between humans and
the climate and the environment; and the incredible complexity of a city
where millions of people live and interact from moment to moment. There
are many kinds of interactions, such as business transactions, bus journeys,
crimes, school classes, car crashes, and chats between neighbours. Yet
simple predictable social behaviour does sometimes arise. For example,
many diverse properties of cities from patent production and personal
income to pedestrians’ walking speed are approximated by functions of
population size (Bettencourt et al. 2007).
The next section introduces the main question of this book and how to
answer it. First, we repeat the core claims above.
The Truisms of Complexity Science
Truisms state the obvious. The following statements will not be obvious to
everyone, but they will be to those working in complexity science.
However, these truisms have not yet been stated clearly and explicitly. They
are the starting point for the analysis of this book because they state the
basic facts about the subject while being compatible with the very wide
range of views about the nature of complexity science and complex systems
found in the literature.
1. More is different.
2. Nonliving systems can generate order.
3. Complexity can come from simplicity.
4. Coordinated behaviour does not require an overall controller.
5. Complex systems are often modelled as networks or information
processing systems.
6. There are various kinds of invariance and forms of universal
behaviour in complex systems
7. Complexity science is computational and probabilistic.
8. Complexity science involves multiple disciplines.
9. There is a difference between the order that complex systems produce
and the order of the complex systems themselves.
The truisms are all independent of each other. Number 6 is an important
discovery of complexity science. Note also that numbers 5, 7 and 8 are not
about complex systems themselves but the science that studies them.
Numbers 6 and 9 are the least obvious and most in need of the articulation
and argument given for them in Chapters 3 and 4.
1.1 What Is a Complex System?
Despite the lack of consensus about how to define complex systems and
complexity, there is a core set of complex systems that are widely discussed
throughout the literature. Chapter 2 presents some of these canonical
examples of complex systems and highlights some of their distinctive and
interesting characteristics. Then Chapter 3 discusses the concepts which are
ubiquitous in the scientific literature about complexity and complex
systems. Ten features associated with complex systems are identified. A
distinction is made between the first four, which are conditions for
complexity to arise, and the rest, which are the results of these conditions
and indicative of various kinds of complexity. The examples of Chapter 2
are considered in an analysis of which features are necessary and sufficient
for which kinds of complexity and complex system. The features are as
follows:
1. Numerosity: complex systems involve many interactions among many
components.
2. Disorder and diversity: the interactions in a complex system are not
coordinated or controlled centrally, and the components may differ.
3. Feedback: the interactions in complex systems are iterated so that
there is feedback from previous interactions on a time scale relevant
to the system’s emergent dynamics.
4. Non-equilibrium: complex systems are open to the environment and
are often driven by something external.
5. Spontaneous order and self-organisation: complex systems exhibit
structure and order that arises out of the interactions among their
parts.
6. Nonlinearity: complex systems exhibit nonlinear dependence on
parameters or external drivers.
7. Robustness: the structure and function of complex systems is stable
under relevant perturbations.
8. Nested structure and modularity: there may be multiple scales of
structure, clustering and specialisation of function in complex
systems.
9. History and memory: complex systems often require a very long
history to exist and often store information about history.
10. Adaptive behaviour: complex systems are often able to modify their
behaviour depending on the state of the environment and the
predictions they make about it.
Some people argue that no scientific concept is useful unless it can be
measured. Many putative ‘measures of complexity’ have been proposed in
the literature, and we review some of the most prominent in Chapter 4 (the
Appendix summarises some of the mathematics used in these measures).
We argue that none of them measure complexity as such, but they do
measure various features of complex systems. We give examples of
measures of almost all of the features of complexity listed above.
Chapter 5 considers complexity science in a wider philosophical and
social context, summarising what we have learned and reflecting on it. We
say what we think complex systems are, argue for our view, and draw
consequences from it. We argue that a system is complex if it has some or
all of spontaneous order and self-organisation, nonlinear behaviour,
robustness, history and memory, nested structure and modularity, and
adaptive behaviour. These features arise from the combination of the
properties of numerosity, disorder and diversity, feedback and non-
equilibrium. We argue that there are different kinds of complex system,
because some systems exhibit some but not all of the features.
We argue that our review of the scientific literature shows that the ideas
of complexity and complex systems are useful in the sense of aiding
successful science. We distill what it is about complex systems that makes
them hard to put in the language of the traditional disciplines and what can
be gained in developing a new language for them. This language allows
descriptions and prediction of complex systems and their behaviour and
features that would otherwise be impossible. The complex systems
discussed in this book, such as beehives, brains and the climate can be
remarkably resilient, but they can also be very sensitive to disruption.
Understanding them is vital for our survival. The final section of this
chapter briefly reviews the history of complexity science.
1.2 A Brief History of Complexity Science
The Scientific Revolution of the sixteenth and seventeenth centuries
involved many crucial developments in science, including the development
of calculus and Newtonian physics. However, experimental and
mathematical sciences are almost as old as human civilisation. Models of
the motions of the heavenly bodies and the behaviour of physical systems
on Earth that were predictive and quantitative existed long before the
development of modern science. Our biology and physics incorporate the
work of Aristotle and Archimedes, respectively, and, while much of
chemical knowledge dates from after modern chemistry developed from
alchemy in the seventeenth century, many basic chemical reactions were
known to the ancients and to Arab and Chinese scholars. Babylonian
astronomers worked out that the Morning Star and the Evening Star are the
same heavenly body (Venus), and much medical knowledge derives from
ancient and medieval research. However, complexity science is very recent,
and this is no accident because, as truisms 5, 7 and 8 above make clear,
complexity science needs a lot of other science.
By the beginning of the twentieth century, physics and chemistry were
highly sophisticated and interconnected, and the theory of probability and
advanced statistical methods were under development. The origins of
complexity science lie in cybernetics and systems theory, both of which
began in the 1950s. Complexity science is related to dynamical systems
theory, which matured in the 1970s, and to the study of cellular automata,
which were invented at the end of the 1940s. By then computer science had
become established as a new scientific discipline. Computation is needed
for anything but the most elementary examination of what happens when a
very large number of parts and their interactions form the system of interest.
There is much more about all this in the rest of this book. In what follows,
the main fields that led to complexity science are briefly described and their
history sketched.
1.2.1 Cybernetics and Systems Theory
The idea that systems in different scientific disciplines follow similar
principles, and that models for one might be useful for others, was
formulated in the theory of cybernetics. Norbert Wiener and Arturo
Rosenblueth coined the term ‘cybernetics’ in 1947 as the theory of control
and communication in living and nonliving systems. Wiener, an American
electrical engineer from MIT, and Rosenblueth, a Mexican physiologist, at
Harvard at the time, were intrigued by the similarities between the systems
they were studying. In particular, the role played by feedback in systems as
different as engineered control systems and the nervous system caught their
attention. Their collaboration began as regular informal meetings of a group
of scientists organised by Rosenblueth at Harvard in the 1930s to discuss
the possibilities of interdisciplinary collaborations. In 1943, Rosenblueth,
Wiener and their colleague Julian Bigelow published a first paper on the
commonalities of engineered and biological systems, introducing the idea
that biological systems use feedback in adaptation (Rosenblueth et al.
1943). Also during this period, Wiener developed ideas similar to Claude
Shannon’s (discussed in Chapter 4) about a mathematical theory of
communication. In 1947, Wiener published his seminal book, Cybernetics:
Or Control and Communication in the Animal and the Machine (Wiener
1961). The notions of feedback and self-organisation, now central to the
study of complex systems (and to much of what follows), were already
present in his and Rosenblueth’s work.
‘Systems theory’ is an overarching term for many fields, including
control theory, cybernetics, systems biology and the study of adaptive
systems. It goes back to Ludwig von Bertalanffy’s work and his 1968 book
General System Theory: Foundations, Development, Applications (1969).
Bertalanffy, an Austrian biologist based in Vienna, wanted to generalise the
observations he had made on biological systems to systems in other
sciences. His early work on a mathematical model of an organism’s growth
over time was published in 1934. Bertalanffy’s goal throughout was to find
models, universal principles and laws applying to systems in general,
irrespective of their particular kind. His achievement was to offer a new
paradigm for conducting science, more than a universal theory applicable to
all sciences (something that is arguably also true of complexity science, as
discussed in what follows). The histories of cybernetics and systems theory
demonstrate that the ground for complex systems science was already
prepared in the mid-twentieth century by the pioneers Wiener, Rosenblueth,
Bertalanffy and others.
1.2.2 Dynamical Systems Theory
The role of dynamical systems theory in the development of complexity
science is various. It is no coincidence that the two fields matured in the
second half of the twentieth century since they both rely heavily on
computation for visualisation and exploration. The most obvious link
between them is that many complex systems are dynamical systems in the
most general sense, and many of them are now modeled using the tools
from dynamical systems theory.
The mathematical sciences aim to predict, accurately and precisely, the
observable behaviour of the world. Predictions can be accurate even if they
are limited in precision. For example, astronomers use models of the solar
system to predict the future positions of heavenly bodies in the night sky
and the dates and times at which eclipses will occur. Their current
predictions can be accurate to the nearest second. Such predictions are
based on observations of the current state combined with laws to calculate
the future state. There is always some error in the measurements that give
the initial data, but in many circumstances this results in a proportionate
amount of error in the predictions. For example, if one calculates the path of
a projectile using Newton’s laws of motion, errors in the magnitude and
direction of the force that launches it and its mass lead to proportionate
errors in the estimation of the position at which it will land. However, the
equations that govern many natural systems are nonlinear in the sense that
the output is not simply proportional to the input. More precisely, a linear
function f is one such that:
f (x + y) = f (x) + f (y) (superposition),
f (cx) = c f (x), where c is any constant (homogeneity).
A nonlinear function is any function that fulfills only one or none of
these conditions.4
Consider population growth. Suppose that every pair of individuals
produces four offspring. Then, in accordance with the superposition
principle above, 200 individuals will produce 400 offspring. In this case,
the function describing it is linear (f (x) = 2x), and a large population grows
in just the same way as a small one. However, real population growth is not
like this because, in large populations, overcrowding causes the growth rate
to decline by increasing the death rate. In this case the function describing it
is nonlinear. An example is the nonlinear equation of the logistic map, f (x)
= rx(1 − x) with parameter r, a simple model of population dynamics and
now a canonical example of a chaotic system. The nonlinearity of
population growth comes about because of feedback. This gives rise to
sensitive dependence on initial conditions by magnifying the effect of small
changes in the initial conditions as the equation describing the system is
iterated over and over again.
Chaos theory is the branch of dynamical systems theory that deals with
systems whose time evolution is highly sensitive to initial conditions.
Chaotic systems appear to be random, because their precise behaviour in
the long run is unpredictable. However, chaotic systems are deterministic in
the sense that their future state is completely fixed by the laws that govern
them and their present state (Strogatz 2014). The notion of sensitivity to
initial conditions is famously given poetic expression by the title of a
presentation by Edward Lorenz at a meeting of the American Association
for the Advancement of Science, ‘Predictability: Does the Flap of a
Butterfly’s Wings in Brazil Set Off a Tornado in Texas?’ (Lorenz 1972).
In general, the longer the time evolution under consideration, the greater
the uncertainty in prediction. In the case of the weather, it is in general
predictable about a week in advance at most. Chaos theory began with
Henri Poincaré’s work in mechanics on the motion of three massive bodies
in accordance with Newton’s laws of motion. He observed that a simple
system can be deterministic but unpredictable in practice because of
sensitivity to initial conditions. Chaotic systems are described either by
differential equations or by the iteration of simple mathematical formulas,
meaning that the result of the calculation is substituted back into the
formula, and this process is repeated many times. Remarkable progress was
made with paper and pencil, but only with the advent of the digital
computer was it possible to perform the vast number of calculations
necessary to study chaotic behaviour in depth and to generate the now-
familiar images of chaotic systems such as the Lorenz attractor.
Many nonlinear and chaotic phenomena were discovered when chaos
theory had matured and complexity science was still in its infancy.
However, chaotic systems are not the same as complex systems. The
relationship between complexity and chaos and nonlinearity is discussed in
Chapters 3 and 4.
1.2.3 Cellular Automata
Cellular automata were originally conceived by Stanislaw Ulam, a Polish-
American mathematician, and John von Neumann, a Hungarian-American
mathematician, in the late 1940s and early 1950s (Ulam 1952; von
Neumann 1966). Von Neumann’s goal in particular, which he achieved, was
to find computational models that could mimic biological self-reproduction.
Cellular automata are arrays of cells in one, two, or more dimensions,
equipped with rules for the cell states and how they can change. The
simplest cellular automata are one-dimensional; they consist of one row of
cells where each cell is in one of two states (for example, one of two
colours). The row of cells is initiated in some configuration of states, and
the configuration is changed in discrete time steps according to a set of
fixed rules. Common to all rules is that the state of each cell at the next time
step depends on its current state and that of its neighbours. Depending on
the exact rules, different dynamics arise. When the rows of cells, each row
representing one time point, are plotted underneath as a two-dimensional
grid, the structures arising can be surprisingly nontrivial.
Two-dimensional cellular automata include the so-called Game of Life
(Guy and Conway 1982). The state of each cell at the next time step
depends on all its eight surrounding cells. Four very simple update rules
lead to nontrivial dynamics that become apparent when the consecutive
configurations are displayed as a video. The name derives from the lifelike
structures that seem to be forming, moving, and disappearing in the video
visualisation. The Game of Life is a now famous example of the emergence
of structure through simple rules and multiple interactions. It can behave
chaotically but can also produce relatively stable patterns with names like
‘eaters’ and ‘gliders’, which capture how they behave over time (‘eaters’
appear to swallow up cells, and ‘gliders’ appear to move across the grid). In
the 1980s the study of cellular automata gained momentum due to the
increasing availability of the computational power needed to simulate them.
1.2.4 The Rise of Complexity Science
One of the first meetings that discussed the connection between cellular
automata, dynamical systems theory, physics, biology, and chemistry was
organised by Doyne Farmer, Tommaso Toffoli, and Stephen Wolfram in
1983 at Los Alamos National Laboratory in the United States (Wolfram
1984). Many of the phenomena that are now often associated with complex
systems were observed in various models of cellular automata presented at
this meeting, including dissipation, phase transitions, self-organisation, and
fractal structure. Participants included James Crutchfield, Peter
Grassberger, Stuart Kauffman, Christopher Langton, and Norman
Margolus.
Around the same time a group of Los Alamos scientists, which included
George Cowan, a former White House science advisor, decided to found a
new kind of PhD-granting institution focusing on interdisciplinary science.
This did not come to fruition due to lack of funding, but the group did
found a research institute in Santa Fe, just down the road from Los Alamos,
in 1984. The first of two founding workshops had amongst the participants
Phil Anderson, Charles Bennett, Felix Browder, Jack Cowan, Manfred
Eigen, Marcus Feldman, Hans Frauenfelder, Murray Gell-Mann, David
Pines, Ted Puck, Gian-Carlo Rota, Alwyn Scott, Jerome Singer, Frank
Wilczek, and Stephen Wolfram, in addition to various scientists from Los
Alamos and from other scientific institutions (Pines 2019). The title of the
first workshop was ‘A Response to the Challenge of Emerging Syntheses in
Science: A New Kind of Research and Teaching Institution’, and it
involved chemists, evolutionary biologists, psychologists and
anthropologists. Transcripts of this workshop were published in 2019 (Pines
2019). In one of the discussions the physicist Norman Ramsey points out
that “almost always in such discussions [of interdisciplinary subjects] you
omit the oldest and most fruitful ones [syntheses], such as between physics
and chemistry”.5
It is noteworthy that ‘complex system’ or ‘complexity’ were not
mentioned among the topics to be discussed. The systems of interest were
taken to be those that had the potential to be the subject of new syntheses
between the sciences, and emergent behaviour was considered a unifying
theme (in line with the emphasis placed on emergence throughout this
book). Examples of what was discussed include the mathematics of
evolutionary theory, war from an evolutionary perspective, brain
mechanisms of hallucination, and modern archaeology.6
‘Evolution, Games, and Learning: Models for Adaptation in Machines
and Nature’ was the title of a workshop held in 1985, again at Los Alamos
National Laboratory, organised by Doyne Farmer and Norman Packard. The
aim was ‘attempts at synthesis rather than reduction’, to address the
unknowns of issues such as the principles underlying evolution and the
operation of the brain using tools from computational and dynamical
systems theory (Farmer and Packard 1986). The list of speakers includes
Michael Conrad, John Holland, Bernardo Huberman, Stephen Kauffman,
Christopher Langton, John Maynard Smith, Norman Packard, Alan
Perelsen, Peter Schuster, and Stanislaw Ulam. Many of the phenomena that
are now often associated with complex systems were reported on at this
meeting, including evolutionary game theory and evolutionary computation,
immune system and machine learning, and neural network models of
learning.
These ‘syntheses of ideas’ all happened around the Los Alamos Lab and
were driven by the people working there and their colleagues in the United
States and Europe. The Santa Fe institute, which has grown in size over the
years, was the first institute explicitly dedicated to the study of complex
systems. It is now one of many research institutes on complex systems
around the world, with most countries in Europe, the Americas, and many
in South and East Asia running at least one. The international Complex
Systems Society was launched in 2004 on a European level. It became
intercontinental in 2006. The next chapter considers a representative sample
of the kinds of system that are studied in these research institutes.
1The term used in much of the literature is ‘complex adaptive behaviour’, but we drop the word
‘complex’ . A similar point was made by Murray Gell-Mann (1994, p. 27).
2For a discussion of emergence in physics see Butterfield (2011a,b).
3Strevens (2016) discusses the relationship between complexity and simplicity.
4If x and y are real numbers, these conditions are equivalent, and if they are real vectors, then
superposition implies homogeneity, but not vice versa.
5To this David Pines replies, “Those are emerged syntheses”. The question he wanted to discuss
was, “Which fields in the natural and social sciences and humanities are the ones where a synthesis is
or might soon be emerging?”
6The initial funding request to the MacArthur Foundation suggested three broad areas of inquiry:
neurophysics, consciousness, and basic physics and mathematics. The funding request was
unsuccessful. Nevertheless, George Cowan and his colleagues were able to raise enough money to
found in 1984 a private research institute located in an old monastery in Santa Fe.
Chapter 2
Examples of Complex Systems
This chapter surveys some of the most common examples of complex
systems studied by complexity science. The first two sections consider
examples from physics and chemistry that show that even nonliving
systems without goals or functions have properties with a rich structure and
can exhibit different kinds of self-organisation and generate order. The first
section is particularly important for the rest of the book, because it explains
concepts that originate in physical science but which have been applied
throughout complexity science (for example, the ideas of equilibrium and
phase transition are widely used outside of thermal physics, where they
originated). The next section is about the universe and how it and its parts,
like the solar system, display various features of complex systems. The final
nonliving system considered is the Earth’s climate, the physics and
chemistry of which both support and are generated by life. The subsequent
sections discuss living systems, which display adaptive behaviour of many
different kinds and different degrees of sophistication, and two complex
systems of human construction – namely, the economy and the World Wide
Web. The chapter ends with the human brain, which is a supreme example
of a complex system. The next chapter identifies the features of complex
systems displayed by these examples.
2.1 Matter and Radiation
Physics is concerned with processes involving matter and radiation at all
scales, from the interactions of subatomic particles and fields to the
formation of galaxies and even the universe itself. At the beginning of
modern physics, in the seventeenth century, some natural philosophers,
such as René Descartes and Pierre Gassendi, revived the ancient idea that
the natural world is based on matter moving around in space. They became
known as the ‘mechanical philosophers’. Descartes thought that space is
completely full of matter, but most others believed in the vacuum,
especially after the invention of the air pump. The original atomists,
Leucippus, Democritus and Epicurus, imagined that everything is made of
tiny particles moving in a void, but they did not have a precise
mathematical account of their collisions and motions. Building on the work
of Kepler, Galileo, Descartes and others, Newton formulated his laws of
motion and the inverse square law of gravitation, from which could be
derived incredibly accurate predictions of the motions of planets, the moon,
and other heavenly bodies. He could also derive the law of the pendulum,
the parabolic motion of projectiles and Galileo’s law of free fall. The
mathematical sophistication of his work set the standard for future physics
and began the era of mathematical science. The predictions of the inverse-
square law of gravitation applied to the motions of the planets are accurate
to one part in 106, even though in the seventeenth century experimental
observations were only accurate to one part in 103. It was not until the
nineteenth century that the orbit of Mercury was measured precisely enough
for better physics to be empirically testable (Penrose 2004, p. 390).
For the Newtonian, the world consists both of particles and forces, such
as gravity and magnetism, that push and pull these particles around. This is
a departure from the mechanical philosophy whose advocates distrusted the
mysterious idea of force. However, there were too many successful novel
predictions, such as the bulging of the Earth at the equator and the return of
Halley’s comet, for such concerns to prevent its widespread adoption. It
seemed that all matter might be just collections of particles and that all the
phenomena we observe arise from the forces they exert on each other and
their myriad motions. However, in the mid-nineteenth century, Maxwell’s
theory of electromagnetism incorporated fields, which came to be thought
of as just as real as particles. Electromagnetic radiation and electronics are
fundamental to everyday technology, such as mobile phones, microwaves
and radar. Our best current physics describes the behaviour and interactions
of matter and radiation in terms of quantum fields that notoriously have
very strange properties and that are not at all like little bits of ordinary
matter. Furthermore, the Newtonian view of space and time as fixed
backgrounds to physical events has been replaced by General Relativity,
which describes gravity in terms of a spacetime that is curved and
dynamical.
However, classical physics is still applicable to much of what occurs in
nature, even in some surprising ways. For example, colliding galaxies can
be modelled as if they were gases, with stars being analogous to molecules,
and the nucleus of an atom can be modelled as a liquid drop (see, for
example, Rohlf 1994, Section 11.3). The statistical methods developed to
describe the vast numbers of particles in macroscopic samples of matter
have been applied to the statistical properties of completely different kinds
of systems, such as neural networks, and ideas from statistical physics are
widely applied across complexity science in many different ways (as
explained in Chapter 4).
Physics is now a vast subject, and only a small part of it is concerned
with the most fundamental theories of strings and quantum gravity. Most
physicists work with models of matter and radiation that describe
approximations or lower energy limits of more fundamental physics. One of
the most important things we have learned from the history of science is
that new, more empirically accurate theories can be related to the old ones
they succeed and the latter retain some validity; for example, ray optics is
an approximation to wave optics, and Newton’s equations are low energy
limits of the equations of relativity and quantum theory. This means that
classical physics has not been replaced for many applications; for example,
fluid mechanics is used to describe what appear to be continuous media
such as flowing water, even though water is composed of molecules and is
far from homogeneous at the microscopic scale. Physics is largely like
chemistry and the rest of science in being about emergent phenomena (in
the sense discussed in Chapter 1 and further discussed below). In what
follows, it is explained how even the physics and chemistry of nonliving
systems exemplify the truisms of complexity science stated in Chapter 1.
As noted in Chapter 1, the starting point of complexity science is the fact
that some of the behaviour of large collections can be novel, in the sense
that the parts on their own, or in small numbers with small numbers of
interactions, do not display it. Emergence is surprising because what will
happen cannot be anticipated by thinking about the behaviour of isolated
individuals or collections involving only small numbers of interactions
among individuals. Most of physics is about the entities, properties and
processes that spontaneously arise when there are very many interactions
among the parts of a composite system. There are many forms that this
emergence can take. The physical world exhibits rich forms of structure at
many different length and time scales, and different physical theories
describe different kinds of emergent behaviour. For example, as discussed
in Chapter 1, a large collection of molecules forms a gas that has emergent
properties of pressure and temperature. There are laws that relate these
quantities, and whole systems can be described using them much more
economically than by describing each of their parts.
In practice, describing the individual behaviour of all the particles of a
gas is not feasible, but it is possible to describe their statistical properties –
for example, the way they are distributed across states with different
energies given the overall temperature of the gas. Statistical mechanics
relates the average motions and energies of the very many molecules in a
gas to its thermodynamic behaviour, such as heat flows and changes of
temperature. This shows how many molecules together can behave in
interesting and novel ways. The study of how macroscopic properties of
solids and liquids arise from the basic physics that describes their
constituent parts is condensed matter physics. It is largely this field that
Anderson drew upon in the celebrated paper ‘More Is Different’, mentioned
at the start of Chapter 1.1
The above examples all involve a higher level of description that
simplifies the underlying very complicated behaviour of a system. The
theory of ideal gases is an extreme example, since it uses just three degrees
of freedom – namely, pressure, volume and temperature – to describe a
system that really has of the order of 1023 degrees of freedom. In general,
emergence in matter is the emergence of order of various kinds. The ideal
gas laws represent a kind of order in the space of possible states of a gas,
ruling out combinations of values of the macroscopic properties that violate
them. They are laws of ‘co-existence’ that apply at any given moment of
time and say what possibilities for different properties are mutually
compatible. Another law of co-existence is the Pauli Exclusion Principle,
which in its simplest form says that no two electrons can have all the same
quantum numbers.2
Dynamical laws represent a different kind of order. Newton’s second law
says how a system acted on by a force changes over time. The Schrödinger
equation in quantum mechanics is a dynamical law, as are Maxwell’s
equations that describe how changing electric and magnetic fields interact
with charges and electric currents. There are many more dynamical laws
describing emergent structure of various kinds. Most take the form of
differential equations. As mentioned in Chapter 1, emergence often means
there are lower-dimensional equations of motion that describe the dynamics
of emergent properties, and the full extent of the dynamics of the system
can be largely ignored. The emergent properties are often the ones we
observe. For example, in the solar system, the myriad gravitational and
other interactions of the very many particles in the planets and the sun give
rise to the emergent order of elliptical orbits of the planets described by
Kepler’s three laws. This is because over large distances only the
gravitational force is relevant, and because the sun is so much more massive
than the planets, so their effects on each other can be ignored. This is
another example of how simple laws can arise. The price that is paid for this
simplicity is that we lose track of the detailed movements of the parts of the
system and only know about their overall behaviour. Furthermore, the orbits
of the planets are not exactly elliptical. The emergent simple laws are
approximate.
The separation of the dynamics of a system into the relatively slow and
fast is one of the most important kinds of approximation in physics (and is
very important for complexity science in general). For example, in a
benzene molecule there are six carbon nuclei and many electrons, but the
dynamics of the electrons can be treated as if it were separate from the
dynamics of the nuclei. This is because electrons are very much less
massive than protons and neutrons, so their motions and interactions are
very much faster. The separation of the time scales of the dynamics is the
basis of the Born-Oppenheimer approximation used in quantum chemistry.
There are many examples of slow and fast dynamics within a single system
in physics, and many other examples in complexity science, and the
importance of this fact is one of the main themes of this book (see Chapter
4, Section 4.3).
Emergence also often takes the form of relatively sharp boundaries in
time, space or both. A good example of a spatial boundary is the
thermocline that occurs in bodies of water where the temperature drops off
sharply so that there are two very different regions rather than a gradual
range. A more complex example is that of snowflakes. They are untold
billions of water molecules forming ice crystals of intricate geometry and of
such variety that the probability of two identical ones forming in the entire
history of the universe is negligibly small. There are about a hundred
different features of large snowflakes, and the number of possible
snowflake structures is of the order of 10100. Yet they all share the feature
of a star-shaped sharp boundary in space. Similarly, there are sharp changes
in time. Dramatic changes in time in emergent structures can result from the
numerous tiny changes in their parts. Consider a snow field on the side of a
mountain. Friction between the individual flakes holds them at some
average depth across the mountain-side. However, gravity is pulling the
snow downhill. A strong gust of wind or a step in the wrong direction can
suddenly cause millions of tons of snow to move at over a hundred miles an
hour.3
Sharp boundaries, and emergent entities and properties in general, are
relative to scales of energy, space and time. At the very small scale, there is
no sharp change in the properties of molecules in a body of water to
correspond to the thermocline, but at larger length scales, such as that
relevant for a diver, the change is stark. Similarly, the dramatic changes
studied in geology take place over time scales of a million or so years, but
over very short time scales everything changes smoothly. In general, there
are time scales over which things change and time scales over which they
do not, and length scales over which they look the same and length scales
over which they do not. It is remarkable that some things look the same at
different length scales or even at all scales (as with a truly random walk or
any fractal structure). Moreover, some phenomena are independent of scale.
For example, the diffusion of molecules in solution and avalanches occur at
all scales in sand and snow. (‘Scale invariance’ is an important concept in
complexity science and is discussed more below and in Chapter 4.)
Gases have very little structure at any scale, but liquids like water are
more interesting, and, when ice crystals form, there are correlations
between the orientation of water molecules over distances about a billion
times the distance between molecules. In this context, order, structure or
form is inhomogeneity of some kind. For example, crystals exhibit order
because the molecules in them are in geometrical structures like the
tetrahedrons of table salt. In the solid state the molecules are not distributed
randomly, but when salt is dissolved in water, the sodium and chlorine ions
are not bonded and the molecules are distributed randomly. Similarly, the
symmetry of snow crystals is actually a reduction in the complete
translational and rotational symmetry of the positions of molecules in liquid
water. The order of ice and snow crystals breaks the symmetry of liquid
water. For this reason, the emergence of order in physics is often
characterised as a process of ‘symmetry breaking’. Much of Anderson’s
discussion of emergence is about how symmetry breaking is found at
different levels in the different sciences. The idea of symmetry breaking is
very general and can be applied to transformations of many kinds, including
very abstract ones. For example, a network of people would be completely
symmetrical if everyone were connected to everyone else so that swapping
people did not change the structure of the network. This amount of
symmetry is never present in a social network because of clustering (see
Section 2.6 and Section 4.8 below).
Symmetry breaking in physics is associated with changes in energy and
temperature and is described by the theories of statistical mechanics and
thermodynamics. Gases turning into liquids and liquids into solids, and vice
versa, are processes called ‘phase transitions’, and they have many special
features. They give rise to relatively sharp changes in otherwise smooth
variation. The transition of water to ice is like this. The structure changes
radically as the temperature changes hardly at all, and ‘latent heat’ is
emitted. Correspondingly, when ice melts, there is a point where adding
more heat does not increase its temperature but is consumed in breaking
down structure. Above the transition temperature, there is no crystal lattice
of any significant size, and below it there is one.4 This abrupt change in
properties is an emergent feature only of large collections. There is nothing
in the basic physics of a small number of water molecules to suggest that
phase transitions would exist.
Large collections of water molecules have a rich structure to their
behaviour, and they also create form in the world. The combination of water
and gravity carves the intricate fractal structure of mountain valleys and
river deltas, producing persistent and large-scale forms, by processes such
as freezing and thawing occurring many, many times. More is different
partly because many iterations is different. Phase transitions of water are
extremely important for life on Earth, but of course there are phase
transitions between solid, liquid and gas in all other kinds of matter too.
Phase transitions are emergent order in the dynamics of systems of many
parts, and they may also produce structure in the systems with which they
interact.5
The sharp boundary in the freezing of water comes where the density
(the ratio of mass to volume) changes discontinuously. As mentioned
above, this involves the emission of latent heat at this transition
temperature. Phase transitions like this, involving latent heat and a
discontinuity in the rate of change of some property of the system, are
called ‘first-order phase transitions’. Density is related to the
thermodynamic quantity of free energy. In thermodynamics, the free energy
is roughly a measure of the capacity of a system to do work and is related to
other quantities, such as the entropy, the temperature, and the internal
energy (more of all this in Chapter 4). The inverse of the rate of change of
free energy with respect to pressure is proportional to density. In most first-
order phase transitions, there is a discontinuity in the first derivative of the
free energy with respect to some thermodynamic variable (in this case the
temperature). However, the idea of phase transitions is much more general
than this even in physics. In ‘continuous’ phase transitions in condensed
matter physics, there is no discontinuous change in the obvious properties
of the system. An example of such a continuous phase transition is
magnetisation. As a lump of iron is heated, its magnetisation does not
change sharply, and there is no latent heat released, but the rate of change of
magnetisation with respect to temperature does change discontinuously.6
There are also continuous phase transitions involving very different
phenomena, including Bose-Einstein condensation and superconductivity.
Phase transitions and broken symmetries can be described by an ‘order
parameter’, which, in the case of the phase transitions of water, is the
difference in the densities of the liquid and gas phases at some pressure and
temperature. This quantity vanishes at temperatures higher than the
transition temperature and gets bigger as the temperature gets lower. The
order parameter for magnets is magnetisation, and there are others for
superconductivity and other phases of matter. An important emergent
feature is the critical point. With water, this is the temperature and pressure
at which the properties of the gas and the liquid are the same and the
transition between them is smooth. Other phase transitions also have critical
points, and associated with them are ‘critical phenomena’ of various kinds.
The study of phase transitions, critical points and the thermodynamic limit
(when the number of particles becomes large) in statistical mechanics gave
rise to a theoretical framework that has subsequently been applied widely
beyond the original context of solids, liquids, gases, and their properties,
like pressure and temperature. Near critical points, fluctuations occur at all
scales. Hence, scale invariance is associated with phase transitions. A
‘universality class’ of models is one in which all the models behave in the
same way near to phase transitions, and hence have the same ‘critical
exponents’ describing this behaviour.
The critical exponent associated with a phase transition features in a
‘power law’, in which one quantity scales as another quantity taken to the
power of the ‘critical exponent’. An example is the degree of correlation
between parts of a large system scaling as a power of the distances from
each other. This occurs when a ferromagnet is cooled down below the Curie
temperature and long-range correlations between the spins begin to form.
These universal aspects to critical phenomena are found in many different
contexts and in many very different kinds of process. Signatures of
universal exponents and power laws are also found in many non-physical
complex systems, ranging from the human brain to modern cities (more on
power laws in Section 2.5 below and in Chapter 4).
Critical phenomena often involve the way that the degree of correlation
changes between parts of a large system that are at different distances from
each other. This occurs when ice melts and the long-range correlations
referred to above break down. Surprisingly, such long-range correlations are
found in systems which are not driven externally towards a phase transition
but are hovering by themselves close to a critical point. The observation
was first made in a one-dimensional lattice of coupled maps (Keeler and
Farmer 1986). It was later observed in a computer simulation of a two-
dimensional cellular automaton called the ‘sandpile model’ (Bak et al.
1988). In this model, ‘sand grains’ fall onto grid cells, and, when too many
pile up on one cell, sand pours onto neighbouring cells in avalanche-like
bursts. The cellular automaton rules are probabilistic and very simple. The
size distribution as a function of frequency of the avalanches obeys a power
law. Both the coupled-maps model and the sandpile model exhibit criticality
that is said to be ‘self-organised’, because no parameters have to be tuned in
order to arrive at the critical point, as they do with, for example, the critical
point of ferromagnets. Real piles of sand don’t behave exactly like the
sandpile model, but they do avalanche at some critical angle of slope
depending on the shape of the grains.
When even very simple physical systems are driven by an energy or
matter influx from the outside, forms of self-organised order can emerge.
For example, when a tray of sand is shaken repeatedly, hexagons can
emerge spontaneously. It is also possible to observe the spontaneous
formation of waves in such a system (Ristow 2000). Self-organised
criticality has been suggested to exist in many systems ranging from
earthquakes to brain dynamics (Pruessner 2012). There is much debate
about the nature of self-organised critical phenomena (Watkins et al. 2016).
The famous Belousov-Zhabotinsky reaction (BZ reaction), described in
1958 by Belousov and in 1964 by Zhabotinsky (Madore and Freedman
1983), illustrates self-organisation in a nonliving system (Nicolis and
Prigogine 1977). The reaction occurs in a mixture of various chemicals,
including potassium bromate and various acids. While the system is driven
by an influx of reagents, various geometric structures spontaneously
emerge. There are chaotic oscillations in the pattern of spots of colour
corresponding to the different chemicals that form and decay and reform. In
this and similar cases, the equations describing the reaction are nonlinear.
There are many examples in chemistry of molecular oscillators, and of the
spontaneous generation of spatial structures, such as the so-called ‘self
assembly’ of molecular nano-structures. These forms of emergence are
found in systems that are ‘open’ in the sense that there is interaction
between them and the environment. The Belousov-Zhabotinsky reaction
and others like it occur only in open systems.
Closed systems tend towards equilibrium states of constant temperature,
pressure, chemical concentrations and other system variables. In general,
equilibrium states are states that do not change in some relevant respect
over some relevant time scale. For example, a cup of coffee is said to be at
thermal equilibrium when it has cooled to room temperature, because its
temperature stops changing over a length scale of seconds and minutes.
However, it is not really static, and if left for a few weeks, it would
evaporate completely. There are many other kinds of equilibrium state
depending on the kind of system – for example, the equilibrium state of a
mechanical system in which none of the parts move and the equilibrium
state of a combination of chemicals in which their concentrations remain
constant. (The importance of the general notion of equilibrium is illustrated
in Section 2.5 below.)7 A crystal at room temperature is like a closed
system in that all its molecules have found their most stable configuration
and its temperature is constant. A gas in a container at equilibrium can be
treated as a closed system, as can systems of condensed matter, even though
they are really interacting through gravitation with the rest of the universe
because the effects of it on them are so small.
An open, driven system does not reach a static equilibrium, but it may
have dynamic equilibria. A dynamic equilibrium is a changing state but one
in which the amount of change is constant and some of the system’s
parameters do not change over time. An example of a system in a dynamic
equilibrium is a lake that receives water through precipitation and loses
water through evaporation and infiltration into the underlying soil. While
the water flows in and out of the lake, the amount of flow is constant, and
water inflow and outflow balance each other out. Closed systems exhibit
only the simplest kind of emergence. More interesting kinds of emergence
occur in complex systems that are out of thermodynamic equilibrium or
driven from outside in some way.
All the truisms of complexity science discussed in Chapter 1 are true of
the study of nonliving matter and radiation. The hierarchical structure of
even monatomic matter, from the subatomic scale to the macroscopic,
shows that more is certainly different in many ways. Thermal and statistical
physics is computational (and probabilistic of course), and many of the
methods developed to study nonliving systems are now applied to living
systems. In the ideal gas laws, there is a very simple kind of emergent order.
Much more interesting kinds of order are found in the structure of phase
transitions. Phase transitions and critical phenomena, in substances such as
water, ferromagnets, superconductors and other exotic forms of matter,
show that there are different kinds of invariance and forms of universal
behaviour among nonliving physical systems. Self-organisation shows how
complexity can arise from simplicity (see Prigogine 1978, 1980). The study
of systems like river deltas involves multiple disciplines such as geology,
hydrology and meteorology. Such systems exemplify how the difference
between the system and the order produced by the system may be hard to
disentangle, because there is feedback between the two as the existing
pattern of erosion directs water into the channels already present, thereby
deepening and reinforcing them. The next section considers the universe.
2.2 The Universe
All living complex systems depend on their histories. This is also true of
many, if not all, nonliving things. It is true, for example, of the Grand
Canyon, in which there is feedback between geology and erosion as water
flows down channels created by previous flows of water. It is especially
true of the universe as a whole, which has a history estimated to be around
13.8 billion years long. While it is not known how large the universe is,
there is a region visible from Earth, called the ‘observable universe’, that is
around 46 billion light-years in radius (one light-year is 9.46 × 1015 metres).
Over very large length scales, above 300 million light-years or so, the
standard model of cosmology regards the universe as homogeneous and
isotropic, which is to say structureless. Closer up, things look very different.
The extraordinary progress in astronomy in the last few hundred years has
revealed that the universe has many different kinds of structure at many
different scales and that the structure of the universe looks different at
different times.
The standard model of cosmology posits an early universe that is both
very hot and very small. No ordinary matter existed, only subatomic
particles and radiation, and there were no stars or galaxies. As this early
universe cooled down, the symmetries of high-energy physics broke, and
the various forces separated. These changes are treated as phase transitions.
In the process known as ‘big-bang nucleosynthesis’ hydrogen, helium,
beryllium and lithium were generated. Once ordinary matter and forces had
arisen, what happened over subsequent millennia was the result of nuclear
forces and electromagnetism combining with the effects of gravity. The
universe now consists of three generations of stars, each with successively
higher concentrations of heavier elements formed by nucleosynthesis in
stars. Type III stars are the oldest, and it is found that they contain very little
metallic matter and no elements heavier than iron (which is number 26 in
the periodic table). The production of heavy elements in stars only occurs
late in their lifetimes, and these elements are distributed into the wider
universe when stars explode in the form of supernovae. Later generations of
stars form from the remnants of previous generations and contain more
metals.
This ancestry of stars gave rise to the tremendous diversity we find on
Earth. There are nearly 100 elements forming nearly 5,000 naturally
occurring minerals, and, including synthetic compounds, there are about 60
million known chemicals. The structures we now see in the universe exist
only because of its history, in which the remnants of stars that have
exploded as supernovae provided the ingredients for the next generation of
stars and the next. The smallest grain of sand on Earth would not exist
without the birth and death of the stars that made it.
The basic components of the universe are galaxies consisting of stars and
remnants of old galaxies, gas, dust and dark matter, all bound together by
gravity, and often, like ours, orbiting a supermassive black hole. There can
be anything from 109 to 1014 stars in a galaxy (1010 in ours), and there are
1011 galaxies in the observable universe. Galaxies come in three main types
– namely, elliptical, spiral and irregular – although there are further
variations, and galaxies are extraordinarily diverse in their properties. For
example, around two-thirds of spiral galaxies, including our Milky Way,
have a bar of stars across the middle. Our galaxy is a highly structured
entity having four spiral arms, as well as regions of gas, and a number of
orbiting smaller satellite galaxies. Galaxies attract each other
gravitationally, and feedback between them affects their formation and
evolution and their motions and collisions. An early use of computer
simulation showed that collisions of galaxies, modelled very simply as
Newtonian particles, could create spiral arms (Toomre and Toomre 1972).
Galaxies collide over time scales of around a billion years, and there are
very few stellar collisions, so at small length and time scales, there is no
collision at all.
Galaxies form groups, clusters and superclusters, where the last of these
contain tens of thousands of galaxies. These superclusters form part of even
larger structures called ‘filaments’, one of which is 500 million light-years
long, 200 million high and 15 million light-years deep. Correspondingly
there are also voids in the universe in which there are hardly any galaxies,
stars or matter. These can be on the scale of hundreds of millions of light-
years. The whole observable universe obeys Hubble’s law, according to
which objects in deep extragalactic space are moving away from the Earth
with a velocity that is proportional to their distance away.
Stars are classified according to the spectra of radiation they emit. The
simplest feature is the colour corresponding to the largest peak of the
spectrum. Colours range from blue through yellow and red to dark brown
and are indicative of the temperature of the star, which can range from
around 500 K to above 100,000 K, with most stars between 2,000 K and
50,000 K. Of course stars have different masses and can also be more or
less bright; hence we have stars such as brown dwarfs and red supergiants.
A graph can be plotted of the brightness of stars against their colours, and it
is found that most stars are found in a band in this graph called the ‘main
sequence’. The properties of stars in the main sequence depend only on
their mass. This is an excellent example of emergent order in the
interconnected dynamics and properties of stars.
As well as stars, galaxies contain dust, gas and plasma that are emitted
by stars when they become supernovae – that is, stellar explosions. (On
long time scales supernovae are events since the explosions themselves last
only for seconds and the expanding clouds of radiation they create only for
days or weeks. On shorter time scales they are objects.) Interactions in
galaxies, therefore, involve dust, gas, plasma, planets, stars, black holes,
globular clusters, satellite galaxies and radiation, and, according to the
standard model of cosmology, dark matter. As mentioned above, it is
thought that feedback plays an important role in galaxy formation. The hot
gas from supernovae creates so-called ‘galactic winds’, and these gases cool
and then stars form out of them, starting the whole process again. There are
also thought to be supermassive black holes at the centre of almost all
galaxies. The space between stars, the interstellar medium, includes cosmic
rays and magnetic fields, but it is almost entirely empty of matter. It
contains only about a million atoms per cubic metre but has clouds of
material within it that can be twice as dense, again illustrating the degree to
which even ‘empty’ space is structured. At small length scales, there is no
structure, because the density of astrophysical gas clouds is much less than
that of what would be regarded as a vacuum on Earth.8
The above-mentioned dark matter is thought to hugely outweigh the
visible matter with which we are familiar. Dark matter is posited to underlie
the structure of the universe, forming filaments and vast halos on which the
galaxies sit. Dark matter is necessary to fully explain the motions of matter
within galaxies. However, it is unclear what properties dark matter has and
whether or not it interacts only gravitationally or harbours new physics.
Nothing is currently agreed about the properties of dark matter, except that
it is massive and does not interact electromagnetically (and hence is dark).
One of the structures in the universe that has been studied most is our
solar system with its planets and their moons and the sun at its centre. The
relative stability of the solar system and its various subsystems, also
including asteroids and comets, is an emergent feature of gravitational
mechanics. The solar system is subject to the electromagnetic effects of the
sun’s solar wind, which also interacts with the local interstellar medium.
The solar wind is a complicated structure that has a rich array of interesting
features. It is a stream of charged particles, including protons and electrons,
which are ejected by the sun. It comes in two forms, slow and fast. The fast
solar wind (which is less dense) is emitted by ‘holes’ in the magnetic field
of the sun, and the slow solar wind is emitted largely from its equator. The
solar wind forms large -scale spatial boundaries. The termination shock is
the region in the solar system where the solar wind slows down sharply to
subsonic speeds. The heliopause is the boundary of a kind of
electromagnetic bubble around the sun and all the planets called the
heliosphere. These, like thermoclines in the oceans, are examples of how at
a large scale there can be relatively sharp boundaries that emerge from
underlying smoothness.
The sun itself is remarkably structured and exhibits a striking form of
emergent high-level dynamics. Every eleven years, for example, the
polarisation of the sun’s magnetic field changes. Another form of emergent
universal structure is the solar system’s disk shape, which is a result of its
history as a dust cloud spinning around the sun. Disk shapes are common in
the universe, because they are naturally created in spinning systems as a
result of basic Newtonian dynamics. Within the solar system, the planets
exhibit all manner of structure of their own, including their climates and
weather and the interactions between them and their moons. For example,
the periods of the orbits of three of Jupiter’s moons are in the ratio 1:2:4.
This is an example of ‘emergent resonance’, meaning that their
gravitational effect on each other is regular and periodic, and in this case it
is stable and self-correcting (Murray and Dermott 1999).
The universe contains a large number of components that interact with
each other in a nonlinear way. There is a nesting of emergent structure on
many spatial scales. Each galactic structure represents the history of the
early universe and the symmetry breaking that gave rise to the fundamental
forces and subatomic particles, as well as the more specific history of the
galaxy’s own formation. For example, the spiral structure of some galaxies
records that they were formed by a collision between two parent galaxies,
and the structure of the Earth records the production of heavy elements in
stars before our sun had even formed. As pointed out above, every grain of
sand has a very long history and requires many layers of emergence, each of
which involves dynamics over different length and time scales and many
open thermal systems. There are high-level laws in astrophysics concerning
emergent entities and their properties, such as the main sequence of stars,
the dynamics of galaxy interactions, and the universe as a whole.
The universe seems to have become more and more complicated since
its infancy, and the most intricate structures in the universe of which we
know (which are living systems) seem to be more complex than they have
ever been. Even if this is so, we do not know if this increase in structure and
complexity will continue or whether the universe will undo the hierarchy of
complexity it has formed. There is also an apparent paradox in supposing
that structure increases as the universe gets older, since it is usually thought
that the initial state of the universe has very low entropy, which in physics
is associated with high levels of order. While the entropy of the universe
has increased monotonically with time, structure and complexity also seem
to have monotonically increased with time over the history of the universe.
Some of the solution of this paradox might be found in the fact that there is
no agreed upon general definition of the entropy of the gravitational field
and so much that is unknown about the universe. Also current thinking is
that everything will end up in black holes, which will evaporate eventually,
leaving the universe in a uniform state and wiping out its rich structure
discussed above. How complexity relates to entropy and disorder and how,
if at all, complexity can be quantified is clarified in the rest of this book,
especially in Chapter 4.
2.3 The Climate System
The Earth’s climate system is at the interface between the living and the
nonliving world. The Earth is 4.5 billion years old. During at least the first 2
billion years of its existence its atmosphere was almost completely devoid
of oxygen (Bekker et al. 2004). The production of atmospheric oxygen was
entirely driven by microbes, which facilitated and coevolved with
geochemical cycles on a surface equipped with only a thin film of liquid
water (Falkowski et al. 2008). The climate and life on Earth are a mutual
product of each other.
It is important to distinguish between climate and weather. Weather is
the condition at a certain point in time and space with respect to
temperature, precipitation, wind and other meteorological parameters.
Climate is, roughly speaking, weather averaged over a time scale of several
decades. Climate is also the statistics associated with weather observations
over time, such as frequency, persistence, and magnitude of hurricanes;
droughts; and extreme temperatures.
The climate system of today is the sum of five major components and
the interactions between them: the atmosphere (air), the hydrosphere
(water), the cryosphere (ice), the lithosphere (land) and the biosphere
(organisms). It is a thermodynamically open system driven by solar
irradiation and also by plate tectonics and mantle dynamics. It changes over
time under the influence of these external drivers, as well its own internal
dynamics (IPCC 2013 2013c; Maslin 2013). It is also affected by the
changing composition of the ocean and the atmosphere due to biochemical
processes and, more recently, human impact (IPCC 2013 2013b).
Although they are open driven systems, the weather and climate systems
exhibit macroscopic patterns which are stable in time, such as the eye at the
centre of a storm or the annually recurring Atlantic hurricane season.
Important large-scale stable patterns in the climate system include the three
different zones of atmospheric convection, which are responsible for the
Earth’s desert belts and the high precipitation zones in-between. The
convection zones originate in the constant variation in solar power influx
with latitude, with time of day and with time of year. This variation is due
to the curvature of the Earth’s surface, the Earth’s rotation around itself and
around the sun, and the tilt of its axis relative to the plane of rotation around
the sun. The result is a temperature imbalance between the equator and the
North and South Poles that leads to physical forces directing air away from
the equator.
The first of the three convection zones, situated between the equator and
30° latitude, is caused by hot air rising at the equator and drifting
northwards and southward until about 30° north and south. Here, the air has
lost most of its moisture and heat and drops back to the surface, where it
forms the first desert belt of the Earth, of which the Sahara is part. At the
surface, the air spreads out both towards the poles and back towards the
equator. At the equator, the air heats up again to close the cycle of the first
convection zone. The air which is not moving back to the equator but
towards the poles meets the colder air coming from the poles at about 60°
latitude north and south. Here, the warmer air is forced upwards by the
colder, heavier polar air. This upward movement creates the first large
precipitation zone and the boundary between the second and third
atmospheric convection zones. These three convection zones constitute a
very simple structure that arises out of the many complicated physical and
chemical interactions between solar radiation, air, land surface and water.
Large-scale stable patterns are also found in the oceans. The greatest
existing currents of water, the deep-water currents, are caused by the
interaction between winds in the lower atmosphere and ocean surface water.
Wind moves surface water via friction. The movement of surface water, as
well as density differences between the layers, causes the movement of
lower-lying water layers. This downward effect sets in motion a circular
heat transport between the northern and southern hemispheres, the so-called
thermohaline circulation, also known as the conveyor belt. The conveyor
belt begins in the Gulf of Mexico, where strong solar irradiation causes
water to evaporate, leaving the surface water very warm and very salty. The
wind currents drive this water north, causing a surface current known as the
Gulf Stream, which transports a hundred times more water than the
Amazon River, the largest river in the world (Baringer and Larsen 2001).
When this stream of salty warm surface water reaches the northern part of
the Atlantic, around Iceland, it cools down and hence becomes denser. Here
it sinks and is driven southward again, along the coast of Europe and Africa
all the way to the South Pole. Around the coast of Antarctica, it mixes with
the extremely cold deep water from the pole. From here, the current
continues eastward to the south of Australia and eventually connects up
with itself in the Gulf of Mexico. This closes the cycle.
The water circulation in the conveyor belt is responsible for the mild
climate zone of Europe, which would otherwise be much colder given its
location across the border of the second and third polar atmospheric
convection zones. It is feared that human impact will lead to the shutdown
of this thermohaline circulation (Rahmstorf 2000). However, interactions
between the many different components and dynamics on different time
scales make accurate and precise predictions very difficult (Clark et al.
2002). For example, the transport processes in the atmosphere and in the
oceans are on different time scales to each other and also on time scales
much slower than the variation of the driving force of solar irradiation.
Feedback is a ubiquitous driver of the state of the climate system. Two
kinds of feedback can be observed: negative (stabilising) and positive
(destabilising) feedback. An example of negative feedback is the
temperature regulation of land surface. An increase in temperature due to
sunlight leads to surface water evaporation, which causes low-level cloud
formation. Clouds reflect a higher proportion of sunlight than clear air.
Thus, cloud formation reduces the amount of sunlight reaching the surface,
and, as a result, the surface temperature decreases again. An example of
positive feedback is the melting of the polar snow cover. An increase in
temperature leads to snow melting, which reveals the soil and rock
underneath. These have a darker surface than ice and are less reflective.
Hence, the cooling effect of the snow albedo is lost, and the further increase
in temperature causes more snow to melt. Radiative processes involving
absorption and leading to a reduction in snow cover generally happen on a
very fast time scale. Hence, any compensating process preventing collapse
would need to be on a similar time scale.
The time scales of processes involved are crucial for the development of
feedback loops. The carbon cycle is a good example for the role of time
scales in the climate system (IPCC 2013 2013b). Carbon exists in various
molecular compositions in all of the five building blocks of the climate –
the oceans, the atmosphere, land, ice and the biosphere. In the atmosphere,
it is most abundant as CO2, but atmospheric CO2 represents only a tiny
fraction of the carbon in the Earth system, the rest of which is tied up in
reservoirs such as soil, the oceans and rocks. There is a constant turnover of
carbon within and between these different reservoirs, the time scales of
which vary from seconds, in animal respiration, to thousands of years or
even, in the case of rocks, millions of years. The take-up of carbon by
marine microorganisms happens on a time scale of hours. On an average,
CO2 molecules are exchanged between the atmosphere and the Earth
surface every few years. The average time it takes for carbon to flow from
the atmosphere to the oceans and back again is around one thousand years.
Natural uptake of carbon into geological formations through chemical
reactions takes a few hundred thousand years. This last cycle, involving
sediment-bound carbon, is called the slow domain of carbon turnover. The
fast domain of carbon turnover includes all the other more rapid processes
such as photosynthesis and ocean absorption.
Until recently the slow domain of carbon turnover had very little
interaction with the fast domain. Any small amount of carbon flux from the
slow domain to the fast domain through volcanic eruptions and chemical
weathering and erosion was quickly absorbed in the fast domain without
affecting the stability of its cycles. However, since the beginning of the
industrial era, set at 1750, the fast and the slow cycles have been strongly
coupled through fossil fuel extraction from geological reservoirs. Between
1750 and 2011, the atmospheric concentration of CO2 has increased by 40%
(IPCC 2013 2013a). The processes in the fast domain cannot balance out
such a massive release of carbon from the slow to the fast domain. Positive
feedback loops have developed with severe consequences for the overall
climate. The most well-known of these positive feedback loops is linked to
the greenhouse gas effect. An increased amount of CO2 and methane,
another carbon molecule (CH4), in the atmosphere leads to increased
reabsorption of the infrared light emitted by the Earth’s surface. This
reabsorption traps heat inside the atmosphere and perturbs the energy
balance between incoming solar radiation and outgoing terrestrial radiation.
Over the last 100 years, the Earth’s global mean surface air temperature
has increased by about 0.5°C. The increase of global mean surface
temperature by the end of the twnty-first century (2081–2100) relative to
1986–2005 is predicted to be between 0.3° Celsius and 4.8° Celsius,
depending on the details of the considered scenario (Pachauri et al. 2014).
The climate system has interconnected processes on many different
length and time scales and both positive and negative feedback loops. It is
the last complex system considered in this chapter which is, at least in many
parts, nonliving. The next sections consider living complex systems and
those constructed by living systems and the new kinds of emergence they
exhibit. These new kinds of emergence include the maintenance of structure
and function and adaptive behaviour of various kinds such as optimisation,
prediction, decision making and ultimately consciousness and thought.
2.4 Eusocial Insects
Eusocial insects show levels of cooperation and organisation unrivalled in
the animal kingdom, as laid out in the now classic book by Bert Hölldobler
and Edward O. Wilson (2008). Eusociality is displayed in species of ants,
bees, wasps, termites, and aphids (small sap-sucking insects). Eusociality is
recognised by four main characteristics: insects live in groups, they
cooperatively take care of juveniles so that individuals care for brood that is
not their own, not all individuals get to reproduce, and generations overlap.
Some of the advanced eusocial insect species have different morphologies
for reproductive and non-reproductive individuals and even for different
kinds of labour within the non-reproductives.
A colony of euosocial insects is a social organisation whose collective
behaviour is comparable to a multicellular organism in which cells
cooperate to keep the organism as a whole healthy, well-fed, and
procreating, while the probability of individual cell death is much higher
than that of the organism. Colonial insects cooperate to ensure the survival
of the colony, while the survival rate of an individual is much lower than
that of the colony as a whole. For this reason, colonies of eusocial insects
are also called superorganisms (Hölldobler and Wilson 2008).
Eusociality is associated with adaptive group behaviour. Swarming is a
form of adaptive group behaviour that arises from very simple individual
behaviours, often involving the exchange of information between
individuals. The next two subsections give examples of adaptive group
behaviour in eusocial ant and bee species and explain how it is the result of
spontaneous self-organisation.
2.4.1 Ant Colonies
The life of an ant colony begins when a queen mates with a fertile male,
after which she digs herself a hole under ground and starts laying eggs. The
eggs hatch after about a week. This first brood immediately begins to work.
The newborns leave the nest to forage and feed the next brood. They also
extend the nest by adding chambers to it. In the meantime, the queen keeps
laying eggs, moving to deeper and deeper parts of the nest. She does not
need to leave the nest ever again because she mates only once in her life.
She stores sperm inside her body and lays up to several hundred eggs each
day for years. In this way the colony grows until it reaches a typical average
size that varies from species to species. Some have about 50 workers;
others, such as those of leaf-cutter ants, grow to a size of several million.
The survival of a newly founded colony is by no means guaranteed. Up to
90% of new harvester ant colonies die before they are two years old. Once
they pass this age, though, they are often robust enough to survive for 20
years and longer (Gordon 2010).
What determines a colony’s survival is its ability to grow quickly,
because individual workers need to bump into other workers often to be
stimulated to carry out their tasks, and this will happen only if the colony is
large. Army ants, for example, are known for their huge swarm raids in
pursuit of prey. With up to 200 000 virtually blind foragers, they form trail
systems that are up to 20 metres wide and 100 metres long (Franks et al.
1991). An army of this size harvests prey of 40 grams and more each day.
But if a small group of a few hundred ants accidentally gets isolated, it will
go round in a circle until the ants die from starvation (Couzin and Franks
2003).
The life of an ant colony is based on communication. Any encounter
between two ants is a form of communication, either by physical touch or
by exchange of pheromones, which are chemicals secreted by the ants. For
example, each worker carries a specific combination of molecules on the
surface of its body. This molecular cocktail contains colony-specific
pheromones and molecules resulting from an ant’s work, such as molecules
picked up on a forest floor indicating a forager ant. An ant’s antenna acts as
a ‘nose’, detecting the molecules by touching another ant’s back. Thus, ants
recognise the type of worker they encounter by that worker’s ‘scent’.
Pheromones are also released by foragers at regular intervals on their way
back from a food source to the nest. Ants searching for food recognise a
nest mate’s pheromone trail and follow it to the food source. They, too, will
secrete pheromones on their way back to the nest and thus enhance the trail
marking. This positive feedback loop results in hundreds of ants going back
and forth between a food source and their nest, thus producing the ant trails
that are so ubiquitous in nature.
The life of an ant colony is stochastic. Ants react to stimuli only some of
the time, and the probability of a response increases with increasing
stimulus. Patroller ants, for example, which are responsible for checking the
safety of the nest entrance, are the first to leave the nest in the morning.
Their departure is triggered by an increase in temperature at the nest
entrance. However, there is no set temperature value at which patrollers
leave. Rather, the higher the temperature is, the more likely a patroller is to
leave. It is their return to the nest that triggers the forager ants to leave the
nest and start their day’s work. The chance for a forager ant to leave the
nest increases with every encounter of a patroller. If there are too few
patrollers coming into the nest, then the foragers do not leave. If this is the
result of some danger outside the nest, then this is a feature keeping the
colony alive. If it is the result of too small a colony and thus too few
patrollers around in general, it is a destabilising feature. In a small colony,
small fluctuations can be fatal.
An ant colony is a social system governed by division of labour: brood
care, nest maintenance, patrolling and foraging. Task allocation is mostly
determined by demand and opportunity in the environment, both of which
are governed by stochastic rules of behaviour based on interactions between
ants but also by genotypic effects (Schwander et al. 2010).
One environmental factor in task allocation is where inside the nest an
ant happens to be. Young ants are more often working as brood carers
simply because they have just emerged from the pupal case themselves and
are therefore already at the brooding site ready to work. Older ants are more
likely to work as foragers, because they have had more time to venture
further away from the place where they hatched. Tasks are not set for life,
and ants may switch tasks very often. For example, the detection of a
plentiful food source leads to the recruitment of ants from other tasks to
become foragers. Similarly, if a predator has eaten many patrollers, foragers
will switch to work as patrollers the next day. If the nest is blocked by
debris, forager ants will be recruited to maintenance until the nest entrance
is cleared. It is the ants who happen to be nearby who are recruited first.
Allocation of reproductive ability is different. Fertility is determined
genetically, but the mechanism for deciding fertility of a larva is currently
unknown (Klein et al. 2016).
Feedback is another mechanism that governs the behaviour of the
colony. For example, given two food sources of different qualities, such as
differences in distance to the nest, ants will quickly and with great
likelihood settle on the source which is closer to the nest. With only a few
ants out foraging, the bias for one source over another is proportional to the
relative quality of the food sources, mathematically a linear relationship.
Once the number of foragers passes a critical value, the preference for the
more convenient source increases drastically. The bias for one source over
another is now proportional to the power of the relative quality of the food
sources, a nonlinear relationship, and it is one example among many
considered by Deborah Gordon (2010).
The combination of stochasticity and feedback can lead to surprising
phenomena. Army ants are known to build bridges across small gaps, such
as between two sticks or a cut in a leaf, in order to get to a food source.
Army ants are completely blind, and yet they build bridges consisting of
hundreds of ants fully suspended across gaps tens of times larger than an
individual ant. A few ants start to form a bridge at the tip of the obstacle,
thus slightly shortening the path for the others. Once the beginning of a
bridge is formed, the ants quickly move it closer and closer to the optimal
connection between the two trail ends by making it longer and, depending
on demand, wider. Position, length, and width eventually plateau out, and
the final structure is surprisingly close to an optimal trade-off between
length of path and number of ants diverted away from foraging to bridge
building. Individual ants do not have the cognitive capacity to perform such
an optimisation on their own. But by following local probabilistic rules, a
colony collectively arrives at a close to optimal decision (Garnier et al.
2013; Reid et al. 2015).
Ants have memory. A forager ant, for example, is able to remember the
number of times it has crossed a pheromone trail. The memory of an
individual ant lasts for between seconds and days, depending on the species
(Hölldobler and Wilson 2008). Forager ants waiting in the entrance of a
nest for patrollers to pass by remember encounters for about ten seconds.
Some forager ants can remember a repeatedly taken trail for several days.
The collective memory of a colony as a whole, however, is generally much
longer than the memory of an individual ant. While an individual red wood
worker ant has a lifespan of about a year, red wood ant colonies remember
foraging trails for several decades. The information is passed on by older
ants guiding newly hatched ants to individual trees. Thus, the collective
memory of a colony is orders of magnitude longer than that of individual
ants.
Individual ants are capable of basic cognitive information processing
which allows the colony as a whole to make its higher-level decisions. For
example, an individual ant assesses the goodness of a potential new nest site
by approximating its area. An ant visits a site several times, each time
laying a pheromone trail while exploring the site. It recognises its own trail
from the visit before and is able to ‘count’ the number of times it crosses its
own path. The fewer crossings it counts, the bigger it perceives the site to
be. The bigger the site appears to the ant, the more likely it is to advocate it
to a nest mate by guiding it in a tandem walk to the new site. If enough ants
consider the site a good choice, the colony as a whole will eventually make
a collective decision to move there. This protocol for assessing the size of
an area is similar to an algorithm devised in the eighteenth century, also
called the Buffon’s Needle problem (Mallon and Franks 2000). Since the
ants engage in a collective choice between two or more alternatives,
following the collecting and processing of information by individuals, this
is a kind of ‘quorum decision making’.
An ant colony is not a closed system. It may have intricate mutual
dependencies with other living systems. As explained by Deborah Gordon
(2010), some ant species build their nests on a plant and have adapted their
behaviour to the specific physiology of the plant. An Amazonian ant
species, for example, cuts the hairs of the host plant and binds them to a
trap using a fungus which the ants cultivate in the nest. This trap provides
the ants with insect prey and at the same time protects the plant from
invaders. Other species prune the ground around their host plant to aid its
growth, which in turn increases the available space for their nests. Such
mutualism between insect species or between insect and plant species is
very common. In some of these mutualistic relationships there is a third
partner involved, so-called scale insects. In horticulture circles, scales are
usually described as pests. But they can form a crucial link between an ant
colony and its host plant. Scales live off the plant’s sap, which they suck out
with their stylet stuck into the plant. They excrete honeydew, which, if not
removed, would eventually drown the scale. The ants protect the scales by
feeding off their honeydew. They protect the plant by fending off insect
predators. All the while the plant provides the ants with a home and feeds
the scales, which in turn feed the ants. Labelling scales as pests is
overlooking such circular dependency between different organisms which is
abundant in the natural environment.
2.4.2 Honeybee Hives
When a young honeybee queen is ready to fly, she leaves her mother’s nest
to mate with male honeybee drones. She mates with about 10 – 20 of them
before she founds her own hive or replaces her mother. A honeybee hive
can grow to a size of up to 50,000 bees (Mattila and Seeley 2007). The
sperm the queen has collected and keeps in her abdomen is enough for a
lifetime. A queen signals that she is alive and well and that no new queen
should be raised by spreading a particular chemical throughout the nest. A
healthy queen lays about 1,500 eggs each summer day, and for each one she
can decide whether to fertilise it. Fertilised larvae that are fed with specially
nutritious food grow to become fertile females, queens; otherwise they
grow into infertile female worker bees. Unfertilised eggs grow to become
male drones. Drones mostly stay inside the nest, feeding on the honey
produced by the female worker bees, or fly out to look for young queens
with whom to mate. The drones’ only task is to pass on the colony’s genes,
which encode the history of its biological development. Most of the eggs
get fertilised because many more workers are needed than drones. The
female worker bees are responsible for maintenance of the nest, brood care,
foraging, and producing honeycombs and the honey to store therein (Seeley
2010).
A honeybee has individual information-processing capacity that goes
beyond that of ants. An example of this is the dance that honey bees
perform in front or on top of their nest mates to advertise a food source or
an alternative nest site. A dance is a walk in half circles, during which the
bees flap their wings and waggle their bodies, for from a few seconds to a
few minutes. The ‘waggle dance’ has always drawn attention but was not
understood until the 1940s. Karl von Frisch, an Austrian ethologist working
at the University of Munich, discovered that the waggle dance was more
than a random jig of excitement over a food source or nest site. Von Frisch
realised that the dance was a symbolic code for the location and quality of a
food source or nest site. The direction of the straight line of the half circle
relative to the main axis of the hive encodes the direction of the location
relative to the direction of the sun. The duration of the dance encodes the
distance, with approximately one second of dance corresponding to a
distance of 1,000 metres. The number of repeats of a dance signals the
quality of the food source or nest site. Foraging bees that watch the dance
and sometimes walk with the bee during the dance put this symbolic
information into action by flying ‘in a bee line’ to the indicated location.
The location information is accurate enough for them to find the site with
very little searching. The remarkable discovery that bees have a symbolic
language to communicate the location and quality of an object, as well as
other contributions to the understanding of collective animal behaviour,
earned von Frisch a shared Nobel Prize in Physiology or Medicine in 1973.
As with ants, task allocation in a beehive is stochastic and modulated by
feedback (Seeley 2009). The workers in a hive decide which task to attend
to depending on the signal they receive from their direct environment and
from other worker bees. A forager returning to the hive with water, for
example, needs to find a receiver bee to whom to unload the water. Any
difficulty in finding a receiver bee signals an oversupply of water. The
longer the search time for a receiver bee, the less likely the forager is to
perform a waggle dance afterwards to advertise the water site. This
stochastic feedback mechanism regulates the number of foragers going out,
reflecting the need of the hive as a whole, while individual bees act only on
local information. Since the decision is stochastic, it is possible that a
foraging bee finds a receiver bee very quickly although the overall need for
water is low. Such misleading events are very rare. The large size of a hive
acts like an error correction for wrong signals. Bee colonies consist of up to
tens of thousands of bees. The law of large numbers dictates that almost all
bees receive the correct signal. The large size of the system makes its
collective adaptive behaviour robust against the occasional non-beneficial
actions of individual workers.
The temperature regulation of a beehive illustrates why hives are also
called superorganisms, as mentioned at the start of this section. From late
winter to early fall, during the brooding season of a hive, the core of the
nest is kept at an almost constant temperature of around 35° Celsius. During
this time the outside temperature can vary between -30° and +50° Celsius.
The bees regulate this stable temperature level through a negative feedback
mechanism. When the hive becomes too warm, the bees form ventilation
tunnels to let in cooler air. When the hive becomes too cold, they heat up
the air by increasing their individual metabolism, fuelled by the honey they
have produced and stored. Another illustration of the superorganism is a
hive’s defense mechanism against fungi. When a hive’s nest gets invaded
by a fungus, the bees inside the hive start flapping their wings to generate
body heat. This produces a fever state of the hive, heating up the nest just
enough for the fungus to die while the bees stay unharmed. Temperature
regulation in a beehive is an example of self-organised adaptive group
behaviour.
Collective decision making is present in all eusocial insects. We saw
several examples of it for ant colonies. Bees have several mechanisms in
place to make a group choice, such as when to move to a new nest. When a
beehive has grown too big for its nest site, it needs to move or split into two
hives. Initially, scouter bees fly out to locate and inspect potential new nest
sites. Back at the hive, they communicate the location and quality of a site
to fellow scouter bees using the language of the waggle dance. The other
scouts are likely to visit the site if it is advertised to be of high quality. If a
scouter bee follows the directions to a site indicated by the dance of another
bee and returns convinced, it is very likely to perform the same dance and
to attract further bees to inspect the site. Low-quality sites are more likely
to be ignored, and scouts will fly out to search for new candidate sites
instead. This feedback mechanism brings more and more bees to a good-
quality site. Swarms can locate, inspect, and report on a dozen sites in one
afternoon alone. The method for deciding on a site is similar to quorum
decision making exercised by ants. The bees sense how many other bees are
at a candidate site at a given time. Once this number exceeds a threshold of
about 20 – 30 bees, they consider this site to be chosen. They return to the
nest, discourage other bees from dancing for alternative sites by head and
thorax butting, and initiate the rest of the hive members to warm up their
wings in preparation for the move. The head and thorax butting ensures that
only one site is chosen and the decision process does not come to a
standstill when two sites have similarly high quality. It is a negative
feedback mechanism necessary to ensure that a decision is made.
More recently, scientists have found remarkable similarities between the
collective decision making of social insects and decision processes in a
primate brain (Seeley 2010). Social insect colonies and primate brains are
clearly very different systems. They are composed of very different
information-processing units, insects in one case and neurons in the other. A
colony is an independent unit, while a brain needs to be part of a body to be
alive. Their tasks are very different, too. Bees forage for food; brains
initiate a thought or the movement of a limb. On the other hand, each
system is composed of units very limited in their individual information-
processing capacity while, as a whole, achieving a remarkable level of
sophistication. A beehive’s procedure of identifying and moving to a new
nest site is similar to the neural process in a monkey’s brain when it is
initiating an eye movement. Researchers presented captive monkeys with a
screen of moving dots, some moving left, others moving right (Britten et al.
1993). The monkeys were trained to decide in which direction the majority
of dots were moving and to indicate their decision by looking in that
direction. During this decision process the neural activity in the brain areas
responsible for visual perception was recorded. The visual system has
direction-sensitive neurons, responding to movements in one direction only,
either left or right. The firing of directional neurons triggers an increase in
the action potential of higher-level neurons in the visual system which
effectively integrate the signal. Once the action potential of a higher-level
neuron passes a threshold, it fires. A ‘decision’ is made in which direction
the object is considered to move. If competing signals are present, it takes
longer until a higher-level neuron exceeds the threshold since neurons that
integrate opposite directions inhibit each other. This threshold-dependent
decision with negative feedback prevents a standstill in the decision process
and contradicting signals being sent to the motor area. It also guarantees
that even small differences are perceived eventually. This neural process is,
like the collective decision making of ants and bees, an example of quorum
decision making. Like ants and bees, markets and economies can be said to
make decisions. For example, markets decide prices. The next section is
about markets and economies.
2.5 Markets and Economies
The first modern economist was the Scottish philosopher Adam Smith. In
his famous book An Inquiry into the Nature and Causes of the Wealth of
Nations, he described economic agents as selfish individuals who solely
pursue their own happiness and are not interested in promoting the welfare
of the society as a whole (Smith 1776). In his view, society nevertheless
benefits from the sum of such individualistic actions. Smith described this
beneficial effect as not only unintended but unpredictable. The behaviour of
many individuals leads to novel consequences, and, he thought, agents
seeking a beneficial outcome for all would in fact be less able to produce
one. He wrote of an ‘invisible hand’ guiding the economy as a whole, and it
is, according to Smith, due to this invisible hand that a greater good
emerges. He was not implying the existence of actual yet invisible central
control in a free market. Adam Smith’s ‘invisible hand’ is a metaphor for
emergent order without central control.
Smith’s ideas are at the heart of classical economics and are often called
on to show that unregulated markets are beneficial. While the benefits of
free markets are much debated, free markets show many emergent
regularities that arise without an overall controller. The law of supply and
demand is a canonical example of an emergent market regularity. In a free-
market economy the price of a product depends on the number of people
who signal interest in buying it – its demand – and on the availability of the
product – its supply. The law of supply states that if the price of a product
goes up, its supply goes up, because more is produced or sourced. The law
of demand states that if the price of a product goes up, its demand goes
down, because fewer can afford it. The two most famous curves in
economics are those of supply and demand, originally drawn by Alfred
Marshall in his 1890 book ‘Principles of Economics’ (Marshall 1890).
These two curves, plotted in the same graph against the price of a product,
intersect at a single point. This is the point of price equilibrium, where the
price a consumer is willing to pay equals the price at which a producer is
willing to offer it.
The equilibration of prices is observed in daily economic life. For
example, the price of fuel can vary between petrol stations nearby in
location. Only a minority of consumers will go to the trouble of finding and
visiting the cheapest petrol station when the difference in price is small.
However, if the difference in price is large enough, more and more people
will drive to the cheapest station. Other stations will adapt, and the prices
will equilibrate again. Stable price equilibria and a balance of supply and
demand are the result of a negative feedback loop between the decisions of
producers and consumers. Recall from the discussion of matter and
radiation that equilibrium states are states that do not change over relevant
time scales. Prices in a market equilibrium still fluctuate but usually not
over the time it takes to fill the tank.
The theoretical existence of a price equilibrium in a market does not
guarantee that this equilibrium is actually reached. While market prices and
demand move towards the point of price equilibrium, as time passes,
advantages may get entrenched, and the equilibrium may never be reached.
Thus, a theory of market equilibrium with a supply and demand law is
compatible with a very unequal distribution of resources. Where this is very
visible is in the distribution of wealth. The observation of unequal
distributions of wealth goes back to the end of the nineteenth century.
Historical data show that 10% of the populations in Britain and France
owned 90% of the total wealth around the turn of the twentieth century
(Piketty 2014, p. 368). The distribution of wealth was studied in detail by
the Italian Vilfredo Pareto, who began his career as an engineer but later
turned his attention to economics and social science (Pareto 1980). Pareto
analysed taxation data and noted that the number of taxpayers in each
income bracket rapidly decreased with increasing income. The distribution
seemed to decay as a function of the income raised to some negative power,
which is a very rapid decay. The inequality observed by Pareto is captured
in the so-called 80–20 rule: 20% of individuals own 80% of the resources.
This is now also called Pareto’s law. Pareto’s law has been found to hold for
a large variety of statistical data. Eighty percent of links on the World Wide
Web, for example, point to only 15% of web pages (Adamic and Huberman
2000; Barabási and Albert 1999). In 2016, 10% of the Earth’s population
owned 89% of global wealth (Credit Suisse Research Institute 2016). Other
examples are the size and frequency of market crashes, the distribution of
earthquakes, the number of species per mammal genus, the number of
interactions per protein, the word frequencies in the English language, the
intensity of solar flares, and the population sizes of cities (for a review, see
Clauset et al. (2009) and references therein).
Pareto’s law of income is an emergent nonlinearity resulting from a
positive feedback loop between the amount of money owned and the
amount of money earned – ‘the rich get richer’. In a finance-dominated
economy in particular, the likelihood of earning money is positively
correlated with the money already owned because the return on financial
capital is usually much higher than the growth rate of the economy as a
whole (Piketty 2014).
Feedback mechanisms give rise to many different economic phenomena.
Another example is the profit on investment known as ‘return’. Companies
invest in bigger production, because it increases their total profit and
because it drives down the cost of production per item. The return per
investment is large at first, but then gets smaller and smaller upon further
investment until it vanishes and no further investment is profitable. This
phenomenon of decreasing profit per additional investment is called
‘diminishing returns’. Diminishing returns are a form of negative feedback.
The system reaches an equilibrium state, which in this case is a steady
supply and price of a product. Diminishing returns are also considered as a
mechanism for maintaining competition. When one company dominates a
market, it will eventually run into the wall of diminishing returns, which
will allow other companies to catch up.
There are counter-instances to the phenomenon of diminishing returns in
which further investment leads to increasing returns. Increasing returns
made Google the dominant provider of online searches and other online
services. There are many other examples of companies growing because if
many people use their services, even more people will want to use their
services, while they will incur relatively small additional cost in providing
for more costumers. Rather than facing diminishing returns, these
companies will grow faster upon further investment. This phenomenon is
not exclusive to the digital economies in the twenty-first century. Alfred
Marshall already remarked on it when he considered the possibility of
monopolies. Both decreasing and increasing returns are the result of many
interactions and are driven by feedback.
In the economic models discussed so far, the economic agents are very
simple in their behaviour. They decide whether to buy or to sell a product
based on its price or on the profit they would make. The modern version of
Smith’s selfish individual is a consumer who is influenced by only two
factors in a decision of whether to buy a product: how much it costs and
how much its possession is aligned with the consumer’s preference. In
microeconomics a measure of preference is called a utility function. The
formal theory of utility has a long history (Stigler 1950a,b). John von
Neumann and Oskar Morgenstern were the first to cardinalise utility (1947).
They showed that, under certain conditions, rational agents make optimal
decisions by always choosing to maximise their expected utility over all
possible options for consumption.9 The von Neumann–Morgenstern utility
theorem forms the basis of what is now called expected utility theory, a
pillar of standard microeconomic theory.
Since von Neumann and Morgenstern’s work, more psychologically
informed versions of economic agents have been developed. Most notably,
in the 1970s, Daniel Kahneman, Amos Tversky and Vernon Smith
introduced ideas associated with what is now known as ‘behavioural
economics’, for which Kahnemann and Smith received the Nobel Prize in
Economics in 2002 (Tversky died in 1996). Behavioural economics has
been used very successfully to model a plethora of phenomena that can be
understood from a psychological point of view but not from the rational-
agent point of view (Kahneman 2003). The importance of irrational agents
in emerging market phenomena such as crashes had been underestimated
until the arrival of behavioural economics.
Financial Economics
Some of the most rapid developments in economic theory took place in
finance theory from the 1990s onwards. One of the dominant assumptions
in economic modelling at the end of the twentieth century was that all
traders are equally well informed and every trader has access to complete
information. In this world of ‘perfect information’, traders who value stock
options take all past events and all foreseeable future events into account.
Since every trader has complete and equal information, a stock price
quickly approaches an equilibrium price which reflects its ‘true’ value. It is
an equilibrium price, because it does not change until new information
becomes available. Financial markets which are run by rational traders with
perfect information never deviate much from equilibrium. Only small
fluctuations about the ‘true’ value can occur. In this perfect world, a stock
price changes significantly only when new information becomes available.
When that happens, information spreads, traders quickly agree on the new
true value of a stock, and a new equilibrium is reached about which the
stock price fluctuates negligibly. The efficient equilibration of a stock
market due to ‘perfect information’ is known as the ‘efficient market
hypothesis’ (Fama 1991). This hypothesis was developed by Eugene Fama
in the 1970s and earned him a shared Nobel Prize in Economics in 2013,
together with Lars Hansen and Robert Shiller. In an efficient market, the
only reason for a major change in stock price, such as a crash, is a major
piece of new information. Bubbles do not exist. The dot-com bubble, for
example, which burst in 2000, could not be accounted for with efficient
market theory. At the end of the twentieth century it was clear that, while
markets are very efficient information processors, they do not equilibrate,
because the time it takes a financial market to reach equilibrium is longer
than the time between events disrupting the system.
Assuming efficient equilibration of prices has consequences for the
predicted statistics of trade data. In an efficient market, stock prices
randomly fluctuate about the equilibrium price. This makes it impossible to
make a profit because randomly fluctuating prices are unpredictable. In
mathematical terms, this means prices move as if they were random
walkers. A random walk is a movement in a random direction for a random
distance at discrete time steps. Stock prices that follow a random walk are
independent of each other, and the amount of fluctuation is independent of
the value of the stock. A deviation from this random movement of stock
prices about their equilibrium value, in a world of rational traders with
perfect information, is caused by external events alone. The internal
dynamics of such a perfect market are such that any external event is
immediately absorbed into a new equilibrium price. The financial system,
according to this random walk hypothesis, is an open system with no further
structure to its internal dynamics, and prices equilibrate on a much shorter
time scale than that of the external events driving the system.
With the advent of computers and computerised trade, it became possible
to generate and analyse large sets of data and thus to scrutinise some of the
traditional assumptions in financial economics such as the efficient market
and random walk hypotheses (Beinhocker 2006).10 It was quickly found
that real data did not match the statistics predicted by a random walk model.
As early as 1963, in the dawn of the digital age but before the arrival of
computerised trade, the mathematician Benoît Mandelbrot suggested that
the statistics proved the random walk hypothesis incorrect (reprinted in
Mandelbrot 2013). Mandelbrot observed that large movements in prices
(i.e., crashes) are much more common than would be predicted from a
random walk model. His work was largely ignored until decades later, when
standard financial theory was in more trouble. In later work, Mandelbrot
and Richard Hudson (2010) found that not only did financial data exhibit
structure, but they also did so over multiple time scales, from minutes to
months and years.
Beginning in the 1990s, some physicists began applying ideas from
physics such as scaling theory and turbulence to finance and economics.
Amongst the pioneers were Eugene Stanley, Jean-Philippe Bouchaud,
Rosario Mantegna, Yi-Cheng Zhang, Doyne Farmer János Kertész and Imre
Kondor. They and others realised the potential of these methodologies to
explain empirical statistical regularities in prices, such as fat-tailed
distributions (a family of distributions of which power-law distributions are
a member) and to design models including agent-based models of financial
markets and human agents’ behaviour.11 One result of these efforts was that
the limitations of equilibrium models of financial markets became apparent.
While they still had their use, concepts from non-equilibrium physics or, as
Farmer and Geanakoplos (2009) called it, ‘the complex systems viewpoint’,
gained traction. Econophysics is now a recognised field with dedicated
research centres and chairs at universities.
As we saw above in the discussion of matter and radiation, one
fundamental property of matter is that it undergoes sudden and dramatic
changes in structure and properties during a phase transition. If a system is
already close to the point of a phase transition (close to its critical point),
only a slight change in external conditions brings about an abrupt change in
the properties of the system. Didier Sornette (2003) advanced the
hypothesis that stock markets are similar to physical systems of matter. He
showed that financial markets, as well as many other complex systems,
carry the mathematical signatures of phase transitions (Sornette 2002).
According to Sornette, stock market crashes are a sign of self-organised
criticality in financial markets.
Crisis in an Economic System
The field of economics is divided into micro- and macroeconomics. The
decisions made by agents and the prices that emerge from them are the
subject matter of microeconomics. The agents might be individual human
beings, but they might also be firms or other corporate entities.
Macroeconomics studies whole economies and their emergent properties of
growth, inflation and unemployment, as well as the effects of government
policy. The two branches of economics are clearly linked, but their
relationship is not fully understood (Ross 2014) and, thus, often ignored.
An instructive case is the 2008 financial crisis, set off by the collapse of the
Lehman Brothers bank in the United States. The crisis, according to a
common charge, had not been foreseen by economists. Macroeconomists
largely believed the economy was stable and that there was little risk of a
crisis. Nonetheless, financial economists had predicted the crisis in
mortgage markets and had published models of the mechanisms by which
this crisis could be transmitted into the extra-financial economy
(Holmstrom and Tirole 1997). The problem was that the financial and the
macroeconomic models didn’t inform one another, while, in the real world,
the financial economy and the macroeconomy are not isolated from one
another.
Whether the crisis was foreseeable or not, it is clear that the events were
contrary to many of the assumptions in standard economic theory. In
addition to the ones considered in the beginning of this section –
equilibrium, linearity, rationality and perfect information – there was
another standard assumption in finance which failed: that of uncorrelated
risk. The assumption in standard financial models is that the risk of one
investment is not correlated to the risk of another investment. In a system
with nonlinear dependencies on overlapping time and length scales, this
assumption is too simplistic. The almost collapse of much of global finance
in 2008 showed that probabilities of default are not uncorrelated. One of the
triggers of what would become a domino effect was the sudden downturn of
house prices in the United States. This downturn left many people unable to
repay their loans on overvalued houses in a collapsing housing market, and
it led to more and more lenders defaulting.12 As a result, insurers had to pay
out on many insurance policies at the same time – policies on which they,
too, defaulted as they were only partially backed up by available funds
(known as liquidity). The highly interlinked structure of the housing market
with the rest of the economy had been hugely underestimated. A single
perturbation of the system at one end – the default of one lender – could
percolate through the system in an unpredicted way and with unprecedented
speed.
The links between economic agents – banks, mortgage holders, investors
– were due to financial products that were relatively new and had a small
market share at the end of the twentieth century but grew rapidly thereafter.
One such product is a collateralised debt obligation (CDO). Generally
speaking, there is a low return on a low-risk investment and a high return on
a high-risk investment. Since the risk associated with mortgage holders
defaulting is too high for some investors (say, pension funds), and too low
for others (say, hedge funds) there was room for profit in designing a
product with more than one level of risk. To this end, debt obligations were
collected into bundles, called tranches, and each tranche had a different
level of risk and return. Now any kind of investors, risk averse or risk
affine, could buy shares in the tranche which was right for them. The less
risky shares, in the so-called ‘senior tranches’, were guaranteed to be paid
out first but had less return associated with them. The more risky shares,
with higher return from the so-called ‘equity tranches’, were paid out only
once the senior shareholders had been paid. Since CDOs are constructed as
bundles, the links between debtors and creditors are hidden. The trading of
CDOs created an increasingly large layer of hidden links in the global
financial network (Beinhocker 2006).
Another, then relatively new, financial product relevant to the financial
crisis in 2008 is a so-called derivative. Derivatives are a form of insurance.
Any creditor can take out insurance on a loan from a third party such as a
large bank. Should the debtor default on the loan, the arising loss of
investment is compensated for by the insurer. Such an insurance is a credit
default swap, or CDS. A CDS allows for zero-risk investments. The United
States in surance corporation AIG used to be a large provider of CDSs. In
2008, CDSs written by AIG amounted to $400 billion (Davidson 2008).
Investment insurance is not regulated in the same way health or property
insurances are. Because of missing regulations, until 2009 a bank could
give out as many CDSs as it liked without actually holding enough assets
(liquidity) to back up the CDSs it had written (Houman 2009). For exactly
this reason AIG collapsed in 2008 (Paul 2008).
Derivatives generated an unprecedented profit for investors, and their
market grew fast in the decade leading up to the financial crisis. With the
increase in derivatives sold, an unprecedented number of hidden links were
created between financial institutions. Warren Buffett saw the dangers of
these financial instruments quite clearly. In a report from 2002, he notes,
“derivatives are financial weapons of mass destruction, carrying dangers
that, while now latent, are potentially lethal” (Buffett 2002). These hidden
layers of the global financial network, created by new financial products,
facilitated and accelerated the spread of default through the system during
the financial crisis in 2008.
Financial Ecosystems
In the wake of the global financial crisis in 2008, standard models of the
financial economy – systems in equilibrium, consisting of rational,
completely informed agents acting in an uncorrelated way – came to be
viewed with much more skepticism. New paradigms for financial economic
models were needed. Modelling the financial economy as a complex
network of financial institutions linked to one another through loans and
insurances is one such paradigm which has gained much attention
(Battiston et al. 2016). Andrew Haldane, former executive director of
financial stability at the Bank of England, advocated the complex network
view as a much needed new paradigm for financial economics (Haldane
2009). According to Haldane and many others, a complex network view of
the economy allows the transfer of insights from other fields which have
used complex network tools for much longer, such as ecology and
epidemiology.
In ecology, network theory is now a standard tool to study species
relationships. In an ecosystem network the nodes are the species, and the
links are predator-prey relations – who eats whom. In a healthy ecosystem,
there is a balance in the diversity of species and in the relative abundance of
predator vs prey. A feature known to enhance the stability of an ecosystem
is the presence of ‘modularity’ in its network of species (Stouffer and
Bascompte 2011). In a modular network not every species is connected to
every other species, but there are pockets of closely linked species, while
each such pocket is only loosely linked to other pockets. Any perturbation
in parts of the system, such as the extinction of a species, is thus prevented
from disrupting the system as a whole. Modularity and diversity are factors
that create stability (Page 2010). The stability of financial networks is
suspected to depend on similar structural properties (Haldane and May
2011). For example, it is reasonable to assume that the extensive trade of
derivatives broke down the modularity of the financial trading network,
which is believed to be one factor in the global financial crisis. Diversity,
too, decreased in the run-up to the financial crisis. Banks’ balance sheets
and risk management had become increasingly homogeneous (Aymanns et
al. 2018).
According to Haldane, one lesson of the financial crisis of 2008 is that a
modest event in a complex adaptive network under stress can lead to wide
collateral damage (Haldane 2009). The inter-linkage of the network
heightens the impact of shocks and crashes and keeps the system as a whole
away from a stable equilibrium. Haldane concludes that financial
economics could learn many lessons from the pandemic of the severe acute
respiratory syndrome (SARS) in 2002. SARS is an infectious and deadly
flu-like virus found in small mammals; it mutated, enabling it to infect
humans. The speed and scale of the subsequent spread among the human
population was enhanced by the global transportation network, which
increasingly links major cities of the world as if they were neighbourhoods
in a single city. The SARS pandemic began with the first known case of
atypical pneumonia in Guangdong Province, China (World Health
Organization 2013). A few months later the first deaths were reported, and
the disease had spread by air travel via Hong Kong to Vietnam and Canada.
Four months after the first reported abnormal case over 300 people had
been infected worldwide and 10 had died. Four weeks later the number had
risen to over 1,300 infected and 49 people dead world wide. While infection
spread, the number of infected grew exponentially. Six months after the
outbreak, a total of 7,761 probable cases in 28 countries had been reported,
with 623 deaths. The SARS virus was identified and its RNA sequenced six
months after the outbreak began (Marra et al. 2003). Due to international
collaboration in prevention measures and laboratory studies, two months
later the World Health Organization was able to declare the SARS
pandemic contained.
It was intervention by governments that ended both the SARS pandemic
of 2002 and the financial crisis of 2008. It is interesting that the
containment of SARS was helped by the existence of another tightly linked
network – the World Wide Web and recently established web-based systems
trawling for unusual health events (Haymann 2013). The Web is the subject
of the next section.
2.6 The World Wide Web
In the 1980s a group of scientists at CERN, the international high-energy
physics laboratory in Geneva, was looking for a convenient way of sharing
data. Some universities and research institutes in the United States had
already built an infrastructure for sharing data, the Internet, by physically
connecting a handful of computers and servers. Various protocols existed
for data sharing, such as the File Transfer Protocol (ftp). The team of
scientists at CERN, lead by Tim Berners-Lee, added a combination of a
programming language (html), which turns text pages into interlinked
hypertext, and a protocol (http), which interchanges these hypertext or web
pages with servers. This allowed data stored elsewhere to be directly linked
to and retrieved via the servers. The World Wide Web was invented. The
servers still need to be connected up physically, but the web pages are
connected only virtually by means of hypertext links. Thus, the World Wide
Web is two networks in one, the physical structure of the Internet and the
virtual structure of the hypertext links between web pages. In 1993 CERN
decided to place this new technology in the public domain. This decision
was probably crucial for the development of the web technology as we
know it today, which is compatible across platforms and browsers and not
fragmented into proprietary islands (see commentary in Berners-Lee et al.
2010).
The World Wide Web has grown enormously since its invention. In
1999, there were an estimated 1.3 billion web pages, of which roughly 800
million were publicly indexable (Lawrence and Giles 1999).13 In 2017, the
number of indexable pages on the web had reached 46 billion
(Worldwidewebsize 2017). The remainder are web pages that are accessible
only if the precise http address of the site or that of another web page
linking to it is known – so-called non-indexable pages – as well as those in
the ‘dark’ net, which are accessible only with specific software,
configurations, or authorisation, often using non-standard protocols.
The Internet – the network of computers being physically linked through
fiber optic cables or otherwise – has also grown massively over the last four
decades. Estimates for the year 2000 differ between 3 million and 100
million servers (Internet Systems Consortium 2012; Lawrence and Giles
1999). One estimate for 2017 is 1 billion servers (Internet Systems
Consortium 2012). An exact count is difficult because there is no registrar
of servers. To estimate the number of servers, algorithms send artificial
messages and record the computers they pass on their way to their
destination. However, estimations of the actual number of computers from
these individual paths are marred with statistical errors. Mapping the
Internet is similar to mapping a network of tunnels by sending through a
robot, which can take only one tunnel at a time. The chances that it will
miss tunnels and entire regions are considerable. For a reliable estimate
many repeats are needed given the size and constantly changing structure of
the Internet.
The total number of web pages, in the billions, is in stark contrast to the
small number of hypertext links one has to follow to reach any given page
from any other page. The average minimum number of links one has to
follow to reach any page is called the diameter of a network. In 1999 the
diameter of the World Wide Web was estimated at 19 (Albert et al. 1999),
which is a very small number compared to the billion web pages existing at
the time. The result of this small diameter has been termed the small-world
effect. The diameter is so small because a very small collection of web
pages has a very large number of pages linking to them (the number of links
is called their degree). These high-degree pages are the hubs, and they form
the highway, so to speak, of the World Wide Web. Any web page can be
reached quickly by going via these hubs. There are fewer than one hundred
domains with more than one million links from other root domains. A root
domain is the name one needs to buy or register, such as google.com. At the
time of writing, the top three root domains with the highest number of links
were twitter.com, facebook.com, and blogger.com (Moz 2018). The
majority of domains has fewer than 100 links. This is reminiscent of
Pareto’s law and the 80–20 rule (see Section 2.5 above). Mathematically
speaking, a large part of the degree distribution of individual web pages
approximately follows a so-called ‘power law’ (Adamic and Huberman
2000; Barabási and Albert 1999). A distribution follows a power law when
the probability of an event (number of links) is inversely proportional to its
size (number of domains with this number of links) raised to some power
(there are more details about power laws in Section 4.6.3 of Chapter 4).
Networks with a power-law degree distribution are also called scale-free
networks. Although real-world degree distributions do only approximately
decay as a power law, and only across parts of the distribution, they are
often referred to as scale-free networks (Clauset et al. 2009). Such ‘quasi-
scale-free networks’ are ubiquitous in technology as well as nature.
Approximately scale-free degree distributions have been reported for
metabolic networks (Jeong et al. 2000), protein interaction networks (Jeong
et al. 2001), film actor collaboration networks (Barabási and Albert 1999),
scientific collaborations networks (Newman 2003), and food webs (Dunne
et al. 2002; Montoya and Solé 2002).
In addition to its hub structure, the World Wide Web is highly clustered.
Groups of web pages are often highly interlinked while only sparsely linked
between groups (Albert et al. 1999). It has been shown that a scale-free
network that also exhibits a high degree of clustering is indicative of a
hierarchical organisation (Ravasz and Barabási 2003). Small clusters of
nodes organise into increasingly larger groups of clusters of nodes. The
small-world effect of the World Wide Web is due to its hierarchical hub-
structured topology, which it has grown into over the years (Huberman and
Adamic 1999).
The Internet is a physical network while the World Wide Web is a virtual
network, yet they have the same topology as a hierarchically organised
network of clusters with a quasi scale-free degree distribution (Faloutsos et
al. 1999). This is striking since they have very different constraints. The
Internet is a network in physical space, while the World Wide Web is not
embedded in any spatial geometry. It costs much more money to build a
server than to create a web page.
A consequence of the clustered structure of the Internet is robustness
against the random failure of servers. The networked structure means that
almost all servers are connected to more than one other server. Should a
server fail, requests for a web page can be routed through other servers. At
the same time the Internet is very vulnerable to targeted attacks. Robustness
against random failure combined with vulnerability to targeted attacks is a
feature of all networks with a quasi scale-free degree distribution (Albert et
al. 2000).
A new global infrastructure is currently being built by embedding short-
range mobile transceivers into more and more everyday devices such as
mobile phones, watches, refrigerators and cars, enhancing communication
between people and devices and enabling communication between devices
themselves. This Internet of Things is interconnecting physical and virtual
things into communication networks (International Telecommunication
Union 2012). The minimum requirement for devices to be part of the
Internet of Things is their support of processing and communication
capabilities. While tagging objects with radio-frequency identification, for
example – as is done for many food and other consumer items today –
allows for the collection of data in real time, so-called ‘smart’ devices can
do more than that; they can process information and trigger processes on
themselves and on other devices. Generally speaking, physical ‘things’ in
the Internet of Things are those capable of being identified, actuated and
connected. Virtual ‘things’ are those capable of being stored, processed and
accessed. A whole new virtual and physical infrastructure is being created,
with its own data sharing protocols and physical networks. Applications of
the Internet of Things today range from ‘intelligent’ transportation systems
and ‘smart’ power grids to ‘e-health’ and the ‘smart’ home. In 2016, the
number of Internet connected devices was estimated to be somewhere
between 6 billion and 18 billion (Nordrum 2016). The Institute of Electrical
and Electronics Engineers (IEEE) expects that by 2020, 250 million cars
will be connected to the Internet of Things and the number of connected
‘things’ will have reached or exceeded 50 billion (IEEE Communications
Society 2015). This number is of an order of magnitude similar to the size
of some of the larger natural networks and not far from the almost 90
billion neurons in the human brain. However, the connectivity of the human
brain is far denser than that of the Internet of Things, since neurons can
have up to 15,000 connections with other neurons (more of these in the next
section).
2.7 The Human Brain
The human brain has a mass of one to one and a half kilograms and is made
up of neurons as well as other cellular matter. The exact number of neurons
in the human brain is unknown, but a recent estimate is 86 billion (Azevedo
et al. 2009), with each neuron having tens of thousands of connections
(DeFelipe et al. 2002). The most striking fact about the human brain is its
ability to produce consciousness, while of course a single neuron or a small
collection of them produces nothing even resembling an idea or a thought.
It is undisputed that the number of neurons and the way they are connected
is crucial for the cognitive capabilities of any organism. Estimates for the
number of neurons in the common dog are around half a billion (Jardim-
Messeder et al. 2017). The octopus vulgaris has about 500 million neurons,
of which two-thirds are located in the arm nervous system (Young 1963).
The honeybee, which is amongst the most sophisticated insects, has about 1
million neurons (Menzel and Giurfa 2001), and the primitive worm C.
Elegans has 302 neurons (White et al. 1986). The basic chemical and
electrical signalling mechanisms of all these animals are identical. In the
evolutionary history of the brain neurons have diversified and become more
sophisticated, but some of the neurons in the human brain are largely the
same as those in the brain of reptiles, and some of the most basic
mechanisms are found in even more primitive creatures (Anctil 2015).
The first scientist to develop a method for recording images of neurons
was the Italian anatomist Camillo Golgi, around 1900. Santiago Ramón y
Cajal, a Spanish anatomist and often called the father of neuroscience,
discovered Golgi’s technique, improved on it and developed his theory of
the brain as a collection of different kinds of cells (Finger 2005). This
would later be called ‘the neuron doctrine’, and it was opposed to the
prevailing view at the time, the reticular theory, according to which the
brain is a continuous mass with no distinguishable components or modules.
Ramón y Cajal discovered the diversity of morphology of brain cells
(which exceeds that of all other cell types in the human body combined).
His main insight was that a neuron has directionality with input and output
corresponding to electric polarity. He proposed this as the basic mechanism
of information processing. The idea that electrical signalling encodes
information still underpins the current scientific model of the brain, though
the role of chemical signalling is now recognised to be fundamental, and
there may be as yet undiscovered mechanisms. In 1906, Ramón y Cajal and
Golgi shared the Nobel prize for Physiology or Medicine in recognition of
their work on the structure of the nervous system.
The brain is hierarchical in its physical as well as its functional structure
(Nolte and Sundsten 2002). It originates at the top of the spinal cord, which
extends from the brain stem down the backbone, and branches into nerves
that extend throughout the body. The forebrain (which, as the name
suggests, is largely though not entirely located in the anterior part of the
brain), specifically the neocortex, is responsible for the most sophisticated,
conscious behaviour. The hierarchical organisation is a result of the brain’s
evolutionary history. New structures are built on top of ancient ones so that
there are vestiges of the reptile brain and the mammalian brain in the human
brain, and the basic mechanisms by which they work are incorporated. The
relative size of these regions has changed greatly in evolutionary history –
for example, the hindbrain is very small in reptiles and much larger in
mammals. The relationship between body mass and number of neurons in
primates follows an approximate power law (Van Dongen 1998). The most
recently evolved part of the forebrain consists of the basal ganglia and the
enveloping two cerebral hemispheres. There are further structures such as
the thalamus, hypothalamus, hippocampus, and the amygdala. In this way
the brain carries the history of its development within its structure.
To a limited extent it is true to say that the forebrain controls the
midbrain, which controls the hindbrain, but there is lots of influence in the
other direction, so this is a big simplification. The hindbrain, which
provides the connection between the spinal cord and the rest of the brain, is
responsible for life-supporting activity such as heartbeat, breathing, and
swallowing.
The pre-frontal cortex of the human brain (the most anterior part of the
forebrain) has increased in size but more importantly become much more
interconnected in the last 2 million years. More than any other brain
structure, it is the cerebral cortex which makes us human. It contains the
machinery for language, conscious perception, the control of voluntary
movements, and intelligence. Modern imaging technology can visualise the
locations in the brain where motor control, visual images, or even feelings
such as regret are generated (see, for example, Coricelli et al. 2005).
Even during a lifetime, much of how the human brain works is plastic, in
terms of normal development during both childhood and adulthood, as well
as in the sense that regions can be often redeployed if others are damaged.
For example, after a stroke other parts of the brain often take over a
function that was temporarily lost because it was associated with a region of
the brain that was damaged (Ward 2004). This form of adaptive behaviour
contributes to the robustness of the brain’s functionality. The brain as a
whole, of course, constantly produces adaptive behaviour in the external
world.
The information processing in the brain, and its input and output with the
environment via bodily perception and action respectively, depends on
transduction, which is the conversion of energy from one form into another.
For example, the photoreceptor cells in our eyes convert the energy of the
light that falls on them into electrical energy, whereas motor neurons, which
end on muscles, turn electrical energy into mechanical energy.
Neuroscientists now have an understanding in fine detail of how neurons
generate chemical and electrical signals (Purves et al. 2018). The basic
structure of all neurons is that they have input and output regions.
Information, in the form of electrical and chemical signals, flows from the
former to the latter. In the output region of neurons the incoming electric
signal activates a chemical (called a ‘neurotransmitter’) which binds to the
connected neuron and thus translates an electrical signal into to a chemical
signal at the synapse, which is the region of connection. There may be up to
tens of thousands of synapses on the surface of one neuron (DeFelipe et al.
2002). Synapses are regulated by glial cells; however, little is known about
how exactly they contribute to information processing in the brain
(Azevedo et al. 2009).
A neuron fires when there is an action potential. The exact point in time
when a neuron will fire is always subject to some amount of randomness,
because action potentials are subject to thermal (and other forms of) noise.
Interconnections between neurons, in the form of synapses, are developed
and maintained by feedback. By firing, a neuron will excite or inhibit other
neurons, and the effects of those excitations on other neurons can reach it
on a time scale comparable to that of its own dynamics. Repeated firing
strengthens the semi-permanent synaptic connections, establishing memory
and learned behaviour, which are associated with groups of neurons that fire
together in roughly the same pattern each time they are activated.
The brain is part of the wider nervous system, and in both there are many
subsystems and processing tasks, often confined within them. For example,
the photoreceptor cells in the retina are connected to other retinal neurons
that process visual information prior to transmission to the brain.
Information is conveyed to and from the brain by nerves but also by blood
vessels, which contain hormones produced by both the brain and the body.
The resulting signalling network has an unimaginable size: ninety billion
neurons, each with several thousand incoming signals, generating a network
of roughly a hundred trillion interconnections.
The brain shows more than anything that ‘more is different’, because of
the richness of the emergence that results from its activity. Language,
thought and mind are the ultimate cases of emergence. Consciousness, the
first-person perspective, our sense of time passing and our awareness of our
own consciousness are the most elusive objects of scientific understanding.
However, the brain also produces an extraordinary range of highly
sophisticated emergent features, most of which are sub- or unconscious. For
example, reaching out to turn a door handle, face recognition and turning
the head towards a sound are all very high-level processes that result from
the firing of specialised neurons in specialised regions of the brain, and
from the action of neurotransmitters, all interacting with the rest of the body
and the environment.
More recently, a ‘hierarchical predictive coding’ approach to studying
the brain has gained traction (Clark 2013). The assumption of the
framework of hierarchical predictive coding is that the brain executes signal
processing as well as signal prediction. The techniques of probability and
computational learning theory are used to model the brain as a hierarchy of
information-processing levels, each level receiving input signals from the
level below and generating signals for the level above. One of the first
examples of hierarchical predictive coding was a model of the visual cortex
(Rauss et al. 2011). At the lowest level of this model, photoreceptor cells
collect visual input from the environment. This information is propagated
upwards to a higher-level processing unit, where the signals are used to
generate predictions of future signals from the lower-level neurons. These
predictions are compared with the true future signals, and any mismatch is
propagated further up through the hierarchy of neural processing levels. At
these higher levels, a similar information processing takes place, and
revised predictions of future input from lower levels are generated. This
leads to feedback loops between the different levels of information
processing. Some of the levels are initiating actions such as motor neurons
initiating a hand movement. On the one hand, the use of predicted signals
speeds up the reaction time. On the other hand, any prediction error will be
costly for the organism. The assumption of the framework of hierarchical
predictive coding is that the brain has evolved into an organism that
minimises predictive error on all levels of its information-processing
hierarchy.
The above illustrates how the brain has many similarities with other
complex systems, such as numerosity, feedback, and probabilistic
dynamics, leading to distributed decision making and a flexible division of
labour. These similarities can go deep, as seen in the example of quorum
decision making in bee colonies (see Section 2.4.2 above). As noted above,
there are very many neurons and very many interactions between the parts
of the brain; equally, there is a lot of diversity in neurons and in the
structure of their connections. The study of the brain involves both
computational and statistical modelling and many disciplines, including
biochemistry, computer science and physics. The brain is the best example
there is of a system that displays adaptive behaviour. The other complex
systems of which individual people are parts, such as markets and social
groups, and the ones we have collectively created, such as the economy and
the World Wide Web exist only because of the complexity of the brain. The
human brain and the collective products of human brains are the most
complex systems known. The next chapter considers what this chapter as a
whole teaches us about complex systems.
1Of course, molecules themselves are emergent entities since atoms are composed of subatomic
particles, and protons and neutrons are themselves composed of quark and gluon interactions.
Ironically, ‘particle physics’ is all about the interaction of quantum fields that are continuous even
though particles are not. (Also deterministic theories can be approximations to non-deterministic
ones and vice versa.)
2The most general form of Pauli’s law says that collections of indistinguishable fermions
(particles of half-integer spin) must be in antisymmetrised states.
3Such dramatic changes are the subject of the branch of mathematics called catastrophe theory.
4Liquid water contains very small clusters of molecules in crystalline form that form and melt
incredibly rapidly.
5The production of structure by a system is how the environment can have a record of a system,
other than just the myriad tiny changes in the rest of the world that the system has caused.
6Since it is the rate of change of a rate of change that is discontinuous, such phase transitions are
often called ‘second-order’. There are even higher order phase transitions. The differences between
different kinds of phase transitions, and their definition, are subtle and highly technical (see Binney
1992).
7A very important kind of equilibrium in behaviour science is the Nash equilibrium of game
theory, which is a situation in which two or more agents in a game have strategies such that none of
the agents has anything to gain by changing them. The idea of thermodynamic equilibrium is not
directly applicable to complex systems such as the economy and the World Wide Web.
8There are of the order of 1025 molecules per cubic metre in the Earth’s atmosphere near the
surface, about 10 million per cubic metre in a very good vacuum in a laboratory, but only about one
per cubic metre in an interstellar gas (Chambers 2004).
9The expected utility of an outcome is the utility of the outcome multiplied by the probability of
that outcome.
10A good account of the history of economics and complexity is given by Beinhocker (2006).
11For an early, comprehensive review, see Farmer (1999).
12‘To default’ means to fail to meet the obligations on a payment.
13A web page is ‘indexable’ if it can be found by a search engine – i.e., if it is reachable by
following links from major known websites.
Chapter 3
Features of Complex Systems
This chapter uses the representative examples of complex systems
discussed in the last chapter to arrive at a list of the distinctive features of
complex systems. Chapter 1 explained the early history of complexity
science in the 1970s. By the late 1990s the ideas and methods of complexity
science had been developed and disseminated widely. There is a snapshot of
the views of prominent practising complexity scientists in a special issue of
Science on ‘complex systems’ (Science April 1999) devoted to a celebration
of the new science. These ideas of some of the key figures are still
representative of the field and are the starting point for our analysis.
1. “To us, complexity means that we have structure with variations.”
(Goldenfeld and Kadanoff 1999, p. 87)
2. “In one characterization, a complex system is one whose evolution is
very sensitive to initial conditions or to small perturbations, one in
which the number of independent interacting components is large, or
one in which there are multiple pathways by which the system can
evolve. Analytical descriptions of such systems typically require
nonlinear differential equations. A second characterization is more
informal; that is, the system is ‘complicated’ by some subjective
judgement and is not amenable to exact description, analytical or
otherwise.” (Whitesides and Ismagilov 1999, p. 89)
3. “In a general sense, the adjective ‘complex’ describes a system or
component that by design or function or both is difficult to
understand and verify. . . . complexity is determined by such factors
as the number of components and the intricacy of the interfaces
between them, the number and intricacy of conditional branches, the
degree of nesting, and the types of data structures.” (Weng et al.
1999, p. 92)
4. “Complexity theory indicates that large populations of units can self-
organize into aggregations that generate pattern, store information,
and engage in collective decision-making.” (Parrish and Edelstein-
Keshet 1999, p. 99)
5. “Complexity in natural landform patterns is a manifestation of two
key characteristics. Natural patterns form from processes that are
nonlinear, those that modify the properties of the environment in
which they operate or that are strongly coupled; and natural patterns
form in systems that are open, driven from equilibrium by the
exchange of energy, momentum, material, or information across their
boundaries.” (Werner 1999, p. 102)
6. “A complex system is literally one in which there are multiple
interactions between many different components.” (Rind 1999, p.
105)
7. “Common to all studies on complexity are systems with multiple
elements adapting or reacting to the pattern these elements create.”
(Arthur 1999, p. 107)
8. “In recent years the scientific community has coined the rubric
‘complex system’ to describe phenomena, structure, aggregates,
organisms, or problems that share some common theme: (i) They are
inherently complicated or intricate . . . ; (ii) they are rarely completely
deterministic; (iii) mathematical models of the system are usually
complex and involve non-linear, ill-posed, or chaotic behaviour; (iv)
the systems are predisposed to unexpected outcomes (so-called
emergent behaviour).” (Foote 2007, p. 410)
Clearly, these people have very different things to say, not all of which
are compatible. Some of these statements introduce ideas that are essential
to complexity; others are too vague or otherwise unhelpful. For example, 1
may be true, but unless we restrict what we mean by ‘structure’ and
‘variations’ everything in the world will count as a complex system.
Comment 2 asks us to choose between equating complexity science with
chaos and nonlinear dynamics, which we argue below is a mistake; or
equating complexity with having a lot of components, which is too
simplistic; or equating complexity with a system with different possible
histories on the one hand, which is again too simplistic, and a completely
subjective answer to our question on the other, which is tantamount to
giving up.
However, together these statements teach us a lot, as follows: 2, 3 and 4
together show the importance of the ideas of multiplicity, nonlinearity and
interaction and the computational notions of data structures, conditional
branches and information processing that are central to complexity science,
as discussed in Chapter 1 and Chapter 4. Comment 4 also introduces the
important feature of collective decision making which is a part of at least
some complex systems.1 Comment 5 raises the important idea, introduced
in Chapter 2, and discussed below, that complex systems are out of
thermodynamic equilibrium. Comments 6 and 7 emphasise that complexity
arises from many interactions and feedback among components, which is
the first of the truisms of complexity science noted in Chapter 1 and which
was encountered throughout Chapter 2 (it is the first feature discussed
below). Comment 8 emphasises the idea of emergence already emphasised
in the previous chapters and discussed further below.
The next section explores these features and others in more depth. We
distinguish between ‘conditions’ and ‘products’, where the former produce
the latter. Abstracting from the quotations above and reflecting on the
examples in the previous chapter gives us the following list of features
associated with complex systems.
1. Numerosity: complex systems involve many interactions among their
components.
2. Disorder and Diversity: the interactions in a complex system are not
coordinated or controlled centrally, and the components may differ.
3. Feedback: the interactions in complex systems are iterated so that
there is feedback from previous interactions on a time scale relevant
to the system’s emergent dynamics.
4. Non-equilibrium: complex systems are out of thermodynamic
equilibrium with the environment and are often driven by something
external.
The interesting thing about complex systems is that these conditions can
give rise to the following products.
5. Spontaneous order and self-organisation: complex systems exhibit
structure and order that arise out of the interactions among their parts.
6. Nonlinearity: complex systems exhibit nonlinear dependence on
parameters or external drivers.
7. Robustness: the structure and function of complex systems is stable
under relevant perturbations.
8. Nested structure and modularity: there may be multiple scales of
structure, clustering and specialisation of function in complex
systems.
9. History and Memory: complex systems often require a very long
history to exist and often store information about history.
10. Adaptive behaviour: complex systems are often able to modify their
behaviour depending on the state of the environment and the
predictions they make about it.
Not all these features are present in all complex systems. Whenever any
of the products are found in a system, they are the collective result of the
conditions, but not all the products are found in all complex systems. Often
products help produce other products – for example, memory is impossible
without a degree of robustness, and adaptive behaviour can build nested
structure and modularity. The following subsections consider each of them
in turn in more detail and begin to assess whether each is necessary and/or
sufficient for complexity on any or some conceptions of what complex
systems are. Chapter 4 revisits these features and discusses them from a
mathematical point of view.
3.1 Numerosity
Mere numerosity of interactions or parts can produce dramatic differences
in behaviour (Anderson 1972). For example, as mentioned in Chapter 2, a
hundred army ants put down on a flat surface will wander around until they
die of exhaustion, but colonies of a million army ants exhibit ‘collective
intelligence’ (Franks 1989). Most of the examples of complex systems
discussed in Chapter 2 have a great many elements. In gases or systems of
condensed matter there are very large numbers of particles. The numbers of
molecules of gas in the atmosphere, transactions in the global economy, and
connections in the Internet are all not just large but enormous. However,
some complex systems have a relatively small number of components. For
example, even a few organisms can engage in collective motion
(swarming), and although a honeybee colony usually contains about ten
thousand bees, the smallest colonies consist of only fifty or so bees. In the
smallest animal brains there are about ten thousand neurons, while since
close to ten thousand neurons die every day in a healthy human brain
without any noticeable effect, ten thousand is not that many in the case of
the human brain.
Although what is meant by a large number of interactions is vague, such
vagueness is ubiquitous in science. For example, the Correspondence
Principle states that quantum systems consisting of a large number of
particles behave classically, without specifying what exactly ‘large’ means.
Similarly, phase transitions and other critical phenomena arise only when
systems are composed of many parts, where, again, no exact number can be
specified just as there is no exact number of birds needed for a flock. In
general, how much numerosity is needed depends on the system. However,
even in those complex systems with relatively few components there are
many interactions among them to generate the relevant complex behaviour,
and in those with many components, like the brain, the huge numbers of
interactions are also crucial to complex behaviour.
Interaction is the exchange of energy, matter or information (which
always involves the exchange of energy or matter). The mediating
mechanism can be forces, collision or communication. Without interaction,
a system merely forms a ‘soup’ of parts that are independent and have no
means of forming patterns or establishing order. Note that interaction needs
to be direct, not via a third party or a common cause. For example, there are
correlations between the pixels on a screen in a video game of tennis, but
the image of a ball is not really caused to move by the image of a racket
hitting it. There is no genuine interaction between the images at all. Thus,
we require not merely probabilistic dependence but causal dependence.
Locality of interaction is not necessary. Interactions can be channelled
through specialised communication and transportation systems that create
long-range interactions between agents of a financial market; nerve cells
transport chemical signals over long distances. It is important that the idea
of interaction here is not just that of the physical dynamics, but of the
dependence of the states of the elements on each other.
In some complex systems, the components are more or less identical, as
with the individuals in a colony of social insects and physical systems
composed of the same kind of molecules. Their collective behaviour gives
rise to new kinds of law-like behaviour that would never be suspected in
advance. The nature of the individual interactions is no different from the
collective interactions, but when there are enough of them, very different
behaviour results. The similarity of the parts is required for them to be
subject to the same laws. However, while numerosity of interactions among
parts is a necessary condition for any of the products of complex systems
above, not all the parts or interactions have to be of the exact same kind,
and indeed it may be important that they are not (we call this ‘diversity’,
and it is discussed further below and in Chapter 4). For example, the
atmosphere is composed of many similar parts, insofar as there are many
units of small volumes of air that are similar and that interact to give rise to
the weather. Yet, of course, oxygen, nitrogen and other gases are different in
important ways in their interactions and properties, and these differences
give rise to important features of the climate.
Often in complex systems, large ensembles of similar elements at one
level form a higher-level structure that then interacts with other similar
higher-level structures. For example, many cells make up a human body,
many human bodies make up a group, and many groups make up a culture
or society. The large ensembles of similar elements in these complex
systems give rise to nested structure and modularity, as discussed below.
3.2 Disorder and Diversity
Liquid water is more disordered than ice, because the orientations and
positions of the component molecules are much less correlated with each
other in the liquid state than when they form ice crystals. Steam is even
more disordered, because there are no intermolecular bonds at all, whereas
there are lots of intermolecular bonds in liquids, which is why they keep a
constant density and do not expand to fill any volume like gases do.
Numerosity is not sufficient for even the most minimal kind of complexity
that is displayed by any system in which there is a degree of spontaneous
order and organisation, because the interactions must be disordered for the
order and organisation to count as spontaneous. Spontaneous order emerges
as a result of random interactions among parts, rather than being built into
the system or controlled externally. For example, a gas displays
spontaneous order during phase transitions. The order arises from the
aggregate effect of many collisions between particles that are effectively
probabilistically independent of each other because the gas has no spatial
structure.
Correlations of different characteristic length and time scales for
different systems are found throughout physics. For example, liquids
exhibit correlations over medium time scales, whereas gases exhibit
correlations only over very short time scales. On the other hand, solids
exhibit correlations over both large time scales and large length scales.
Where there is order, there is predictability in principle, though perhaps not
in practice. Highly disordered systems can still have predictable statistical
properties. Even gases have some order in the sense that their molecules are
typically distributed among different energy states in a way that depends on
the temperature of the whole gas. A completely disordered system would be
one that is totally random in the sense of lacking any correlations between
parts over space or time at all scales. The idea of complete disorder or
complete order is an abstraction, and any real system has elements of both
order and disorder.
As discussed in Section 2.1 of Chapter 2, order is inhomogeneity of
some kind, and that means with reference to some set of properties, so
disorder ought to mean homogeneity. Yet clearly in one sense a gas is much
less homogeneous than a solid, because the positions of the molecules in a
gas change much more over time. On the other hand, a gas at constant
temperature is very homogeneous in the sense that the average distribution
of its molecules stays the same, and, as pointed out in Chapter 2, there is a
complete rotational and translational symmetry about it. Ideas of order and
disorder are always applied to specific features, and care must be taken to
define clearly what the relevant notions of order and disorder are.
In complex systems, disorder can exist at the lower level in terms of the
stochasticity in the interactions between the parts, as well as at the higher
level, in terms of structure which emerges from them and which is never
perfect. Disorder in the interactions is a property of all the examples of
complex systems discussed in the previous chapter. In many biological
processes order at the higher level emerges from disorder at the lower level,
and thermal fluctuations are necessary for them to take place. For example,
protein binding to DNA is driven by the energy provided by thermal
fluctuations (Sneppen and Zocchi 2005). The result is the transcription of
DNA and the production of new proteins.
The fact that the order at the emergent level is never perfect and disorder
remains has led to the idea that complexity lies between order and disorder
(Waldrup 1992). In fact what lies between order and disorder is the
structure of a complex system or the structure that is produced by a
complex system (see Ladyman et al. 2013).
The absence of centralised control is another kind of disorder.
Centralised control is when a special component controls an aspect of a
system’s behaviour. For example, thermostats control heating systems
centrally, and herds of elephants exhibit coordinated motion because they
follow the dominant female in the group. Many complex systems have to
maintain control over some parameter, often their temperature, as with
beehives and brains, but they are often not equipped with single control
actions. Rather, in complex systems, order, organisation and control are to a
greater or lesser extent, distributed and locally generated, and not centrally
produced. With decentralised control no privileged individual is issuing
commands, and yet parts still display coordinated behaviour, such as the
collective motion of a flock or swarm or the temperature regulation of a
hive. In the economy there are central banks and policymakers, but their
degree of control of the relevant parameters is partial at best. Clearly, lack
of central control is not sufficient for complexity because there may be a
lack of control or order without this producing anything.
Diversity is when the components of a complex system vary, in their
physical structure and perhaps also in their role. Quote 1 above takes such
variation to be essential to complexity, and it is indeed found in many of the
examples discussed in Chapter 2. Many sophisticated forms of adaptive
behaviour would be impossible without diversity, since it facilitates the
division of labour. For example, in brains there are neurons specialised to
different tasks, and in financial markets there are specialised components
such as banks. Diversity is not important to the flocking of birds, however.
Diversity is neither necessary nor sufficient for any form of complexity, but
it is important in many complex systems. Different forms of disorder and
diversity are discussed further in Chapter 4.
3.3 Feedback
The interactions in complex systems are iterated so that there is feedback
from previous interactions, in the sense that the parts of the system interact
with others at later times depending on how they interacted with them at an
earlier time. Feedback plays a crucial role in how disorder generates order,
and it can give rise to stability and robustness. The presence of feedback in
a system is not sufficient for complexity because it does not always give
rise to some kind of higher-level order or stability.
In a sense feedback is present in all systems, if only because their parts
gravitationally interact with each other and because those interactions
depend on the results of earlier interactions, and so on. However, in
complex systems the feedback is relevant to the dynamics of the system as a
whole, and to be so it must take place on a similar time scale. In physics,
most processes happen on well-separated time scales, which is the opposite
of what happens in the domains of the other natural sciences. For example,
feedback is irrelevant to the aggregate behaviour of gases but vital to
chemical oscillators such as the BZ reaction. The climate illustrates well the
role of time scales in feedback. The slow and fast carbon turnovers do not
exhibit feedback; one happens on a time scale of thousands of years, the
other on a time scale of minutes to days. The fast turnover finds a new
dynamic equilibrium much more quickly than the relevant dynamics of the
slow process.
In living systems and some systems that are derivatives of living
systems, feedback can be what produces adaptive behaviour. All examples
of such complex systems in Chapter 2 exhibit feedback, including flocks,
hives, brains, markets and IT networks. Consider again the flock of birds.
Each member of the group takes a course that depends on the proximity and
bearing of the birds around it, but after it adjusts its course, its neighbours
all change their flight plans in response in part to its trajectory; so when a
bird comes to plan its next move, its neighbours’ states now reflect in part
its own earlier behaviour. In flocks and swarms feedback between the parts
that start randomly distributed in their positions and movements result in
correlations in their motions. In this way the order of the whole arises
spontaneously. Feedback is what brings about coordinated behaviour and
maintains the structure of the collective, despite the absence of a central
control system.
Ants are able to undertake complex tasks such as building bridges or
farms even though no individual ant has any idea what it is doing; left on its
own it exhibits much simpler behaviour. Individual ants behave very
differently when enough of them repeatedly interact with each other. The
repeated interactions lead to feedback, and they bring about the building of
nests and the rearing of offspring that is impossible for a small group with
too few interactions.
Recall the distinction between positive and negative feedback made in
Section 2.3 of Chapter 2. Negative feedback can give rise to stability, which
can be a form of error correction, as in the brain, when feedback from
sensory systems is used to correct motor systems, and as in the feedback
between producers and consumers in a market economy. Feedback is often
to be understood as feedback in a functional sense – for example, feedback
of information. Feedback as error correction is often used in control
systems, the paradigm of which is the Watt steam regulator, where the
speed of rotation of the device interacts in a feedback loop with the steam
engine to control the speed of the engine. However, this is central control as
opposed to distributed feedback because there is a single component that
has a special role. In sum, feedback is a necessary but not sufficient
condition for any kind of complexity.
3.4 Non-Equilibrium
In Section 2.1 of Chapter 2 the thermal equilibrium of a system is defined
as the state in which the thermodynamic properties of its macrostate are
effectively unchanging. This is the macrostate that is reached after some
time, whatever the exact initial microstate. The free energy of a system is its
capacity to do work, and this is minimal at thermal equilibrium, while the
entropy is maximal (given the value of conserved quantities such as the
total energy). The thermodynamic equilibrium states of gases and
condensed matter systems and their phase transitions have a rich structure.
Apart from isolated samples of gases and condensed matter systems, the
other examples of complex systems discussed in Section 2.1 are
thermodynamically open.
An open thermodynamic system is a system with a net influx of energy
or matter, which can be thought of as a free-energy gradient (either within a
system or between connected systems). Open systems may be externally
‘driven’, as with a pile of sand being driven by the adding of more sand and
the BZ reaction being driven by a flow of the reagents. The amount of order
of some kind generated by complex systems can often be related to the
steepness of the gradient of some quantity. For example, minerals are
created in the Earth’s crust by processes involving great heat flows and
changes of pressure. More kinds of minerals are produced where there is
greater variation in temperature and/or the concentration of chemicals
(Hazen et al. 2008). Both of these examples can be associated with a steep
gradient of the free energy of the system with respect to time. The dominant
driving force of the Earth’s climate is solar radiation
Systems with a net influx and outflux of matter or energy can be in a
dynamic equilibrium (like the lake of Section 2.1). A system is said to be in
‘dynamic equilibrium’ or ‘steady state’ if some aspect of its behaviour or
state does not change significantly over time – for example, the ratio of
types of elements and their interactions. Many biologists think about
adaptive behaviour as the maintaining of dynamic equilibrium in a
thermodynamically open system. The maintenance of the steady internal
state of an organism is called ‘homeostasis’. Homeostasis more generally is
a form of dynamic equilibrium in which a system that is changing is
nonetheless stable in some respect. The free energy is important in living
systems, because maintenance of structure is associated with the
consumption of free energy (well adapted mechanisms minimise the
amount of free energy needed). Living systems are open systems and
maintain themselves out of thermal equilibrium by consuming food or, in
the case of plants, by photosynthesis.
The notion of dynamic equilibrium applies to systems whose physical
parameters, such as temperature and energy, are irrelevant to their
dynamics, because of the way in which they are driven. For example, the
temperature and energy of a bank’s physical state are irrelevant to how it
functions as part of the economy. A financial market is an example of a
system that can be in a state of dynamic equilibrium and for which a
temperature distribution is not defined. Since many complex systems are
not physical systems, and hence have no specified temperature or free
energy, the concept of thermodynamic equilibrium as such is not applicable
to them. However, because of the tight link between thermodynamics and
statistical mechanics and probability theory (Jaynes 1957), generalisations
of temperature and thermodynamic entropy are used – for example, for the
equilibrium analysis of complex networks (Albert and Barabási 2002).
The BZ reaction, discussed in Section 2.1 of Chapter 2, forms a type of
‘Turing pattern’ (Turing 1952). Alan Turing proposed a reaction-diffusion
model to reproduce how patterns such as stripes can form in an embryo. It
has since been shown that the original model by Turing could explain most
biological self-regulated pattern formation (Meinhardt 1982). In these
examples of pattern formation after a period of time, a steady state is
reached in which the system maintains a pattern of colours, such as spirals,
stripes or spots. During the steady state, chemical reactions are taking place,
while the overall ratio of visible colours and hence molecular constituents
remains approximately constant.
A system is in chemical equilibrium if the concentration of reactants is
constant and if no external source supplies energy or matter to the system.
In a state of chemical equilibrium, only reversible reactions are taking
place, and no net energy is added to the system. The BZ reaction is not in
chemical equilibrium because the concentration of the reagents changes
over time. This notion of equilibrium used in chemical reactions is an
instance of the more general notion of the dynamic equilibrium of a driven
system that is widely applied to complex systems (see Section 4.4).
Numerosity and disorder and diversity are found in an isolated gas at
thermal equilibrium, but as noted above, being thermodynamically open is
necessary for the most minimal kind of complexity displayed by matter. In
general, physical complex systems are out of thermodynamic equilibrium
with the environment and are often driven by something external. These
conditions, together with feedback between the parts of the system, produce
the other features of complex systems (spontaneous order and organisation,
nonlinearity, robustness, history and memory, nested structure and
modularity, and adaptive behaviour) that make them of such importance and
interest.
3.5 Interlude: Emergence
There is no conception of complexity or complex systems that does not
involve emergence. As discussed in the brief history in Chapter 1, the idea
of emergence was central to the discussions that founded the Santa Fe
Institute. The problem is in saying exactly what emergence is over and
above the minimal idea of novelty in the sense of the whole behaving in
ways that the parts in isolation do not.2 As explained in Section 2.1, almost
all of physics and all of chemistry deal with emergent properties in this
sense, as exemplified by all ordinary material substances. For example,
water is a transparent liquid that forms ice when frozen and steam when
boiled, but none of these properties or processes can be applied to a water
molecule or the atomic constituents of water – namely, hydrogen and
oxygen atoms – which are themselves emergent from subatomic particles
and fields. Similarly, a gas exhibits the properties of pressure and
temperature and the gas laws relating them. The kinetic theory of gases,
according to which heat is molecular motion and pressure is due to
molecular collisions, connects the emergent properties of the whole with the
properties of the parts.
Different kinds of emergence exemplified by the examples of the last
chapter are listed below.
• The emergence of structure at different scales
Ultimately as far as we know the physical structures at different scales in
nature emerge from fundamental particles and fields. This is the kind of
emergence exemplified by atoms and molecules, crystals and minerals,
liquids and gases, geological features and the atmosphere of planets, and
solar systems and galaxies. Sometimes the structure is highly ordered
physical structure in space – for example, in the lattice of salt crystals.
Sometimes the structure is in the way the system changes and in its
dynamics, as in the BZ reaction (see Section 2.1).
As emphasised in Chapter 1 (truism 9), it is important to distinguish the
structure and order of a system and the structure and order it produces –
beavers and beaver dams, bees and honeycombs, and elephants and their
graveyards. The last kind of order or structure may persist long after its
production has stopped (human burial mounds are visible millennia
later). Structures produced by complex systems are not in general
complex systems themselves but may be studied as products of them.
• The emergence of dynamics, properties and laws
The gas laws are emergent, and they relate properties of a whole gas that
its parts do not have. One important kind of emergence is that there is a
lower dimensional effective dynamics within a higher dimensional state
space. That is the case for a ball rolling on a table. The number of
degrees of freedom can be massively reduced if the collection of
particles constituting the ball is described as a single entity. Similarly,
the approximately elliptical orbits of Kepler’s laws of planetary motion
emerge over time from the gravitational interaction between all the
particles of the sun and the planets. Emergent properties and laws
massively reduce the number of degrees of freedom needed to describe
and predict the behaviour of the system.
However, there is a trade-off as the emergent laws do not say anything
definite about the individual parts of the system. In general, descriptions
of systems at a higher level say only probabilistic things about the lower-
level entities that make them up, but this is often all that is needed. The
alternative would be that, in practice, nothing could be said at all about
the system dynamics because the calculations would be impossible even
with supercomputers. Compare trying to predict what will happen next
when a mouse and an owl catch sight of each other using the laws of
fundamental physics applied to their parts with using behavioural
biology.
As explained below, emergent dynamics or emergent laws may be
nonlinear even though the underlying system is linear. The higher-level
dynamics may even appear random while the system is deterministic, as
is the case with chaotic systems. Often there is emergent relative
decoupling of dynamics at different time and length scales, as illustrated
by the examples of the climate and the solar wind.
• The emergence of various kinds of invariance and forms of universal
behaviour3
Critical phenomena, phase transitions and scaling laws are another kind
of emergent structure that can be found in very different physical
systems, as discussed in Section 2.1. In complexity science critical
phenomena show up in probabilistic descriptions of geological
phenomena and in network descriptions of biological and social systems,
as discussed in Chapter 4. Recall truism 5 from Chapter 1 which states
that complex systems are often modelled as networks or information-
processing systems. Doing so allows novel kinds of invariance and
forms of universal behaviour to be studied.
(1) Networks
The parts and interactions in complex systems may often be considered
abstractly and independent of their physical nature. This makes complex
systems well suited to being modelled as networks. Complex systems are
composed of many components interacting repeatedly, but some
components may interact more than others, and some may be well
connected and others less so. A new kind of universality emerging from
this point of view is a scale-free degree distribution (see Section 4.6).
Many of the properties and products of complex systems discussed
below can be found in networks. Network theory is discussed in more
detail in Chapter 4 and in the Appendix.
(2) Computation/information-processing systems
Bold claims are sometimes made about how life, or even the universe
itself, performs computations or maybe that somehow it is just
computations or information (see, for example, Wolfram 2002). This
may be wrong, but it is certainly true that it can be extremely useful to
model systems as such. The multiple pathways of quote 2 at the
beginning of this chapter are emergent at a higher level of description
than the physical description of a complex system. They may be
computational pathways or flows of information. The emergent level of
description in which the system is considered as processing information
is of particular importance in complexity science. Many measures of
complexity use information theory, as discussed in Chapter 4.
Probability and statistics are the underlying tools for the study of
networks and information, and the application of statistical mechanics to
complex systems outside of physics, as explained in Chapter 4, relies on
them.
The next sections consider various emergent features of complex
systems that take different forms in the context of the different kinds of
emergence above.
3.6 Order and Self-Organisation
Clearly a fundamental idea in complexity science is that of order and self-
organisation in a system’s structure or behaviour that arise spontaneously
from the aggregate of a very large number of disordered and uncoordinated
interactions between elements. Even simple condensed-matter systems and
gases exhibit the minimal kind of order and self-organisation discussed in
Section 2.1. The dynamic equilibrium in the patterns generated by chemical
reactions like the BZ reaction is a kind of spontaneous order. Spontaneous
order is exploited by applications of complexity science such as methods of
‘self assembly’ in chemistry. The generation of order involves symmetry
breaking. For example, when a crystal forms, the translational and
rotational symmetry of the molecules in the liquid is broken. Self-
organisation may mean collective motion and various kinds of organised
behaviour, as discussed below.
It is far from easy to say what order and structure (as well as self-
organisation) are. Notions that are related include symmetry, periodicity,
determinism and pattern. As pointed out in Section 3.2 above, the notion of
order may mean so many things that it must be carefully qualified if it is to
be of any analytical use in a theory of complex systems. One of the most
difficult issues is how order and self-organisation in complex systems relate
to the information content of states and dynamics construed as information
processing. The problem is that the interpretation of states and processes as
involving information may be argued to be of purely heuristic value and
based on observer-relative notions of information being projected onto the
physical world. Chapter 4 returns to the role of information theory, and the
idea of order is made mathematically precise in Section 4.5 of Chapter 4.
It is a necessary condition for a complex system that it exhibit some kind
of spontaneous order and self-organisation. Of course, order and self-
organisation must be relatively stable to be worth identifying. As with the
notion of dynamic equilibrium, what counts as ‘stable’ depends on the
relevant time and length scales (see the discussion of robustness below). All
the other products of complex systems require some kind of order or
organisation.
3.7 Nonlinearity
In the popular and philosophical literature on complex systems a lot of heat
and little light is often generated by talk of linearity and nonlinearity. For
example, in his widely cited book, Klaus Mainzer claims that “linear
thinking and the belief that the whole is only the sum of its parts are
evidently obsolete” (1994, p. 1). It is not explained what is meant by ‘linear
thinking’ nor what nonlinearity has to do with the denial of reductionism.
Unfortunately the discussion of complexity abounds with non sequiturs
involving nonlinearity. However, nonlinearity must be considered an
important part of the theory of complexity if there is to be one, since
certainly many complex systems are also systems that obey nonlinear
equations such as power laws.
As explained in Chapter 1, the mathematical definition of linearity is that
a system is linear if one can add any two solutions to the equations that
describe it and obtain another, and multiply any solution by any factor and
obtain another. Nonlinearity means that this ‘superposition’ principle does
not apply. It may refer to dynamical equations or to any equations
describing the relationship between variables. Nonlinearity is often
considered to be necessary for complexity (as suggested by quotations 2, 5
and 8 at the start of this chapter). This claim is misleading. Complex
networks, for example, which are much studied in complexity science, can
be defined by matrices which are mathematically linear operations. There
are also complex systems subject to game-theoretic and quantum dynamics,
all of which are subject to linear dynamics. This shows that, in general,
nonlinearity in the dynamics of the elements is not necessary for
complexity, while nonlinearity on the level of the whole is often a feature of
complex systems.
Nonlinearity is also not sufficient for a complex system (on any but the
most all-encompassing conception), because, for example, a simple system
consisting of a pendulum can be subject to nonlinear dynamics and exhibit
chaotic behaviour, but it exhibits only the most minimal kind of emergence
without self-organisation. Complexity is often linked with chaos, and as
noted above, it may be conflated with it, but the behaviour of a chaotic
system is indistinguishable from random behaviour. It is true that there are
systems that exhibit complexity partly in virtue of being chaotic, but their
complexity is something over and above their chaotic nature. Furthermore,
since chaotic behaviour is a special feature of some deterministic systems,
any dynamical system that is stochastic is by definition not chaotic, and yet
complexity scientists study many such systems.4
So it seems that chaos and nonlinearity of the underlying dynamics are
each neither necessary nor sufficient for complexity.5 The fact that
nonlinearity is a commonly mentioned feature of complex systems is an
example of the conflation of having nonlinear dynamics and being
complicated yet defined by simple rules (where ‘complicated’ means
difficult to predict).6
However, there is a sense in which complex systems are never linear
sums of their parts. The emergent behaviour of complex systems comes
about because the parts interact so what the individuals do together is
different from the sum of what each one of them does alone. For example,
individual ants will just wander around, but together they will behave very
differently. Hence, nonlinearity may be both a property of the system’s
dynamics that contributes to its complex behaviour and it may also be a
product of the complex system in the sense of being a form of emergent
order – for example, power laws in statistical distributions of city sizes or
income (see Section 2.5). Nonlinearity in some guise, either of dynamics or
of emergent relations, is a very common feature of complex systems, and it
is discussed much more in Chapter 4.
3.8 Robustness
Robustness is of some kind of structure, order, self-organisation, law or
function. The most basic kind of emergence occurs only when the order that
arises from interactions among parts at a lower level is robust. It should be
noted of course that such robustness is only ever within a particular regime.
The emergence of condensed matter requires a certain range of temperature,
because at very high energies, molecules all break down and atoms lose
their electrons and become plasma. At even higher energies there are no
atoms only subatomic particles, and so on. In living systems, the conditions
in which emergent structure is robust may be very limited. For example, the
interactions among our neurons generate an emergent order of cognition,
but if the brain is heated to about 5° Celsius above its normal temperature, it
breaks down.
The order in complex systems is often robust because, being distributed
and not centrally produced, it is stable under perturbations of the system.
For example, the order observed in a flock of birds, despite the individual
and erratic motions of its members, is stable in the sense that the buffeting
of the system by the wind or the random elimination of some of the
members of the flock does not destroy it. A centrally controlled system, on
the other hand, is vulnerable to the malfunction of key components.
The term ‘resilience’ is sometimes used interchangeably with
‘robustness’, but it is better thought of as a kind of robustness - namely, the
ability to recover from a perturbation on a time scale which is short
compared to the system’s lifetime. Resilience may be formulated in
computational language as the ability of a system to correct errors in its
structure, and this may be achieved by exploiting feedback. In
communication error correction can be achieved by introducing a limited
amount of redundancy. This redundancy is usually not explicit, such as a
copy of the string or its parts, but more subtle – for instance, exploiting
parity checking, which is more computationally intensive but also more
efficient (the message is shorter than simple duplication) (Feynman 1998).
In his influential account of complexity (discussed in Chapter 4) Charles
Bennett (1991) specifically mentions error correction: “Irreversibility seems
to facilitate complex behavior by giving noisy systems the generic ability to
correct errors”. For example, a living cell has the ability to repair itself
(correct errors), as when a malfunctioning component is broken down and
released into the surrounding medium. In the cell, a small perturbation
cannot be allowed to spread to all the degrees of freedom. Hence, the cell
has a one-way direction for this dispersal, errors within the cell are
transported out, and possible sources of error outside the cell are kept out
(assuming the errors are sufficiently small).
Many similar elements interact in a disorderly way in a gas, but it is not
a complex system in any but the most trivial sense of obeying emergent
laws. However, a system consisting of many similar elements that are
interacting in a disordered way has the potential to form patterns or
structures. An example is the Rayleigh-Bénard convection patterns that
form in layers of liquids subjected to a temperature gradient, as well as the
patterns made by the BZ reaction and other self-organising chemical
phenomena. On an appropriate time scale the order is robust. This means
that although the elements continue to interact in a disordered way at small
length and time scales, the larger scale patterns and structures are preserved.
A macroscopic level arises out of microscopic interaction, and it is
relatively stable. This kind of robust order is a necessary condition for a
system to be complex. A good example of robustness is the climatic
structure of the Earth, where rough but relatively stable regularities and
periodicities in the basic phenomena of wind velocity, temperature, pressure
and humidity arise from an underlying nonlinear dynamics.
Some kind of robustness is necessary for any kind of order to exist and is
therefore necessary for complexity. However, order may persist without
being maintained by the kind of emergent processes born of stochastic
behaviour that is characteristic of complex systems. On the other hand, in a
completely random system perturbations do not affect the state. Hence,
robustness is not sufficient for order (or any other products of complex
systems).
There is obviously a trade-off between robustness and adaptation,
because systems that are very robust adapt more slowly. A salt crystal is
very robust until it is dropped on the floor and breaks to pieces. It can
reform but only after being dissolved in water and being left until the water
has evaporated. To conclude, robustness of some kind is a necessary
component of every complex system and is discussed at length in Chapter
4.
3.9 Nested Structure and Modularity
Anderson (1972) argued for the importance of considering hierarchies of
structure to understand complex systems. There is nested structure in the
physical world from the subatomic through the atomic to the chemical, and
ultimately to planets, stars, galaxies, clusters and superclusters. The Earth
and its oceans and atmosphere exhibit a very rich nested structure, as does
the solar system. In such complex systems there is structure, clustering and
feedback at multiple scales.
In systems that perform functions there can be the division of function or
labour and the emergence of levels of functional organisation. The division
of labour in an ant colony is an example of functional organisation, in the
sense that there are subsystems for subtasks, but this is true even though the
worker ants are more or less identical. In many other cases functional
organisation goes along with differentiated structure that is specialised to
fulfill different functions, as in even simple cells. Multicellular organisms
divide labour such as respiration, digestion and excretion, as well as
cognitive capacities such as memory and perception, among highly
specialised parts.
Cities, economies and many other social systems are systems composed
of other complex systems, and often have many levels of organisation that
exhibit what Herbert Simon (1962) called ‘hierarchical organisation’ (the
hierarchy is of system and subsystem). There is an emergent hierarchy in
the life sciences, because living systems and collectives are organised into
different units that have different roles and that can be said to represent the
state of the environment and internal states of the system itself. The
ultimate result of all the features of a complex system above is an entity that
is organised into a variety of levels of structure and properties; these
interact with the level above and below and exhibit law-like and causal
regularities and various kinds of symmetry, order and periodic behaviour.
The best example of this is the whole system of life on Earth. Other systems
that display such organisation include individual organisms, the brain, the
cells of complex organisms and so on. ‘Modularity’ includes the
differentiation of structure within a system, like the clustering of
connections in a network, and also the organisation of the parts of a system
to perform different functions, which often goes along with corresponding
structural modularity (see Chapter 4).
3.10 History and Memory
Complex systems have order and structure that persist, and they also
produce order and structure that persist. Persistence is always over some
relevant time scale. It seems that nothing lasts forever. Eventually plants
and animals die and their bodies break down. Even stars and galaxies have
a finite, albeit very long, existence. The fate of the universe itself is
unknown, but the things within it carry parts of its history with them
because their complex structure is a product of that history and could not
exist without it (as explained in Section 2.2). The Earth, with its heavy
elements, is the product of and carries information about the history of the
evolution of the universe over several generations of galaxies.
Living complex systems evolve over countless iterations of life cycles of
generation and corruption. Genetic inheritance carries part of that history
from parent to offspring. Every multicellular organism contains basic
cellular mechanisms that were features of the most primitive life on Earth.
As discussed in Chapter 4, Bennett proposed logical depth to measure
the history of complex systems. Information theory can readily be used to
quantify history as correlations over time. “A system remembers through
the persistence of internal structure” (Holland 1992). Any robust order that
exists in a system can be thought of as memory, but in general we refer to it
as ‘history’ and reserve the term ‘memory’ for something more specific. It
is not helpful to think of a footprint in the sand as the beach’s memory of a
visitor because the footprint does not cause the beach to do anything
differently. However, if an animal remembers the smell of a relative, this
affects its behaviour. Hence, we use ‘memory’ only in relation to systems
that exhibit adaptive behaviour and have internal degrees of freedom that
are used to represent how the state of the environment or the system itself
was at some time for the purposes of information processing.
3.11 Adaptive Behaviour
Adaptive behaviour is what organisms do to survive and reproduce. It
ranges from the relatively simple and inflexible behaviour of a bacterium
swimming up a chemical gradient in search of nutrients to the most
sophisticated human deliberation. The American complexity scientist John
Holland defines systems with adaptive behaviour as “systems [that] change
and reorganize their component parts to adapt themselves to the problems
posed by their surroundings.” (Holland 1992).
Living complex systems organise themselves to do something. For
example, neurons perform computational tasks, birds flock to migrate and
army ants spontaneously build bridges over obstacles when foraging. So
adaptive behaviour performs some function. Numerous kinds of adaptive
behaviour are studied in biology, and many kinds of adaptive behaviour
occur not just in living systems, but also in those complex systems
derivative of life discussed in Chapter 2. Whether or not it is appropriate to
attribute functions to physical systems that are not either biological or
derivative of biological systems is highly contentious. We reserve the term
‘adaptive behaviour’ for living complex systems and complex systems
produced by living complex systems. Collective motion is one of the main
examples of adaptive behaviour in living systems, but recall that eusocial
insects and other social organisms display many other forms of collectivity,
such as the collective behaviour of bees cooling down their hive when it is
too hot and the collective decision-making about the site of a new nest or
whether or not to tie up ants in building a bridge over an obstacle rather
than going around it. The brain can be modelled as neurons collectively
making decisions.
The complex systems human beings have built all have a purpose or
function. For example, the Internet is used to send information, and markets
are for the exchange of goods and services between individuals or groups.
The adaptive ‘behaviour’ which arises is communication efficiency, and the
feedback between traders’ actions and prices leading to markets finding
stable prices. Adaptation may take the form of optimisation, but results are
usually less than optimal. The collective functioning of all the nodes of the
Internet or other networks, and even evolutionary change in natural or
artificial systems, can all be modelled as emergent adaptive behaviour and
decision making. All the products of complex systems except nonlinearity
have functional versions of various kinds. Functional organisation is the
functional version of order. Robustness often means retaining the structure
that is required to perform some function or maintaining a parameter such
as temperature. Robustness can also be functional robustness in the
presence of structural change, as when the Internet is said to be robust
because traffic is automatically rerouted when a part of the system fails.
Modularity is the functional version of nested structure, as with, for
example, the modularity of the Internet. As noted above, memory is best
thought of as a functional notion. For example, the immune system
remembers how to fight antigens it has encountered before because it has
subsystems whose physical degrees of freedom can be copied to prepare
antibodies. Most human artefacts are made to perform a function, and even
art can be construed as providing for some kind of need. People readily
think in functional terms, and many of the concepts of complexity science,
such as information, computation, organisation and memory are functional.
In fact, it is hard not to slip into functional ways of thinking because they
are so important and so useful. Even the climate can be thought of as having
a function since it provides and maintains an environment for life.
3.12 Different Views of Complexity
There is a long-standing debate about whether or not emergent entities and
properties are real in the way that material things are. For example, are cash
flows real? They might instead be regarded as just emergent levels of
description that should not be taken as describing real things in the way that
physics does. For example, if one plays a video game, one imagines that a
player hitting a ball on the screen is what makes the ball move, but of
course there is no causal interaction between the separate parts of the
screen, only between the circuits in the chips of the computer generating the
image. This may explain why in quote (2) (on Page 63) the idea of a
subjective description is raised. Note, however, that most of physics is
about emergent entities but few would want to say that, for example, the
solar wind is not real. Chapter 5 returns to this issue.
Emergence is often contrasted with reduction, and, as noted in Chapter 1
people often associate complexity science with the failure of reduction. For
example, Anderson (1972) talks about the limitations of reductive methods
in describing systems composed of many parts. Indeed, many treatments of
complex systems come close to defining complexity science in terms of the
rejection of reductionism. For example, Melanie Mitchell (2011) begins her
book Complexity with a discussion of antireductionism, which she defines
as the view that “the whole is more than the sum of its parts.” Yet in the
very same paragraph she goes on to say how complexity science “explain
how complex behaviour can arise from large collections of simpler
components.” This paradox runs throughout the literature on complex
systems. On the one hand, there are calls for reductionism to be rejected,
yet on the other hand, complexity science is all about breaking systems into
parts and explaining the behaviour of the whole that results. The paradox is
resolved by the fact that it is the interactions among the parts that make for
the behaviour of the whole.
There are many different conceptions of complexity. Considering which
of the examples of complex systems in Chapter 2 have which of the features
above allows for several viable conceptions of complexity. The most
inclusive conception of complexity is that it is just emergence, and it is
exhibited in different ways by each of the examples. In fact, it is exhibited
by all matter due to its emergent properties and hierarchical structure from
the subatomic scale upwards. Even a monatomic gas at constant pressure,
temperature and volume exhibits the emergent order of the ideal gas laws,
even though there is no order at the level of the individual molecules. A
more restrictive view that is still very broad is that complexity is self-
organisation, where this is a special case of emergence. This view excludes
isolated systems of matter that only exhibit emergence of the simplest kind
but includes many systems in condensed matter and the universe, as well as
all living systems.
Many of the products of complexity, such as order and organisation,
memory and robustness, as well as some of the conditions, such as disorder
and feedback, are readily described in computational/information-theoretic
language. To understand complex systems it is often possible and indeed
necessary to ignore the details of the physical interactions among the parts.
All that seems to matter are the abstract properties of these interactions, not
exactly how they work. So, for example, the swarming behaviour of insects
can be understood without regard to whether the individual animals are
signalling to each other by scent, sight or sound. The crucial thing is that
they are able to exchange information about their behaviour and the state of
the environment. Complex systems are often thought of as maintaining their
order and producing order by the exchange of information among their
parts. Abstracting systems to networks also facilitates describing them in
computational terms, and (as discussed above) models of very different
systems in terms of their parts and the interconnections between them
exhibit kinds of invariance and forms of universal behaviour. There are
many computation/information-based conceptions of complexity and
complex systems (associated with some of the measures discussed in
Chapter 4).
Numerosity, diversity and disorder, feedback, non-equilibrium and
nonlinearity are all features that can be had by nonliving physical systems
without functions, goals or purposes. On the other hand, adaptive behaviour
is shown only by biological systems or those derivative of them. As noted
in Chapter 1, many conceptions of complexity confine it to systems that
exhibit adaptive behaviour. For example, Mitchell’s (2011) examples of
complex systems are all living systems (insect colonies) or parts of livings
systems (the immune system, the brain) or derivative of living systems
(economies, the World Wide Web) because she takes complex systems to be
restricted to those that adapt. We call this the ‘functional’ view of
complexity, because the idea of adaptation is applicable only to biological
systems or other systems that perform functions to achieve goals or
purposes which are nonliving systems made and sustained by living
systems. Chapter 4 argues that the various measures of complexity in the
literature do not measure complexity as such but measure different forms of
the various features of complex systems explained in this chapter. Chapter 5
returns to the big questions about complexity and which if any of the above
views is correct.
1Recall from the discussion at the end of Chapter 1 that some people think that complexity is
confined to systems that display adaptive behaviour. We return to this issue below.
2As Paul Humphreys says, “In the case of emergence, there are too many different uses of the
term ‘emergence’ that are entrenched across various fields for a single comprehensive definition to
be possible at this point in time or, perhaps, ever.” (2016, p. 26). A detailed analysis of the current
disagreements about emergence is provided in Wilson (2015).
3This is truism 6 from Chapter 1.
4Note that chaos as in ‘chaos theory’ is always deterministic chaos.
5Robert MacKay (2008) argues for a definition of complexity as the study of systems with many
interdependent components and excludes low-dimensional dynamical systems, and hence many
chaotic systems.
6Arguably, unlike complexity, chaos can be simply defined as so-called strong mixing, meaning
that the system’s state is effectively independent of its earlier state (Werndl 2009).
Chapter 4
Measuring Features of Complex
Systems
This chapter is a guide to quantifying complexity based on the fruits of the
analysis of the previous chapters. Many measures of complexity have been
proposed since scientists first began to study complex systems, and the list
is still growing. The main lesson of Chapter 3 is that complexity is a multi-
faceted phenomenon and that complex systems have a variety of features
not all of which are found in all of them. This implies that assigning a
single number to complexity cannot do it justice. As the physicist and
Nobel laureate Murray Gell-Mann noted early on, “A variety of different
measures would be required to capture all our intuitive ideas about what is
meant by complexity” (1995, p. 1).
If complexity is a collection of features rather than a single phenomenon,
then all quantitative measures of complexity can quantify only aspects of
complexity rather than complexity as such. This insight makes it prudent to
ask what any purported ‘measure of complexity’ actually measures. In the
final section of this chapter, a few, by now classic, measures of complexity
from the 1980s and 1990s, mentioned in many discussions on the subject,
are discussed, including effective complexity, effective measure complexity,
statistical complexity, and logical depth. The discussion confirms that they
each quantify one or two of the features identified in Chapter 3.
The fact that complexity measures ever quantify only one or two features
of complexity and never the phenomenon as a whole should inform any
practitioner’s approach to quantifying complexity. The chapter goes through
the features of complex systems identified in Chapter 3 and discusses
mathematical means available to quantify them, accompanied by examples.
A feature can take different forms in different scientific disciplines, and,
therefore, there often exist several ways by which to measure any given
feature. The techniques were often invented in a different discipline from
those in which they are now applied as the provided examples from the
scientific literature illustrate.
This chapter demonstrates the truism of complexity science that it is
computational and probabilistic (truism 7 in Chapter 1). It also further
explains some of the new kinds of invariance and forms of universal
behaviour that emerge when complex systems are modelled as networks
and information-processing systems (truisms 5 and 6). The distinction
between the order that complex systems produce and the complex systems
themselves is central to the analysis (truism 9).
The ‘classic’ measures that are discussed in the final section of this
chapter were constructed as thought experiments rather than as measures to
be applied to real-world systems. Hence, they do not belong in a
practitioner’s tool kit. However, they have played a role in developing our
understanding of complexity over the decades, and they serve now to help
us understand the distinction between measuring complexity and measuring
features of complexity.
This chapter is accessible to a general reader, with the possible exception
of Section 4.10, which contains the discussion of the classic measures of
complexity. It is a more technical section compared to the rest of this
chapter. Throughout this chapter, terminology is explained when it is used
for the first time. Further mathematical background is given in the
Appendix.
4.1 Numerosity
The most basic measure of complexity science is the counting of entities
and of interactions between them. Numerosity is the oldest quantity in the
history of science, and counting is among the most basic scientific methods.
Counting is the foundation of measurement because quantities of everyday
relevance such as length, mass and time can be counted in units such as
metres, grams and seconds. Counting alone does not tell us what counts as
‘more’ in the sense of ‘more is different’, because, as noted in Chapter 3,
when we consider emergent behaviour, how many is ‘enough’ depends on
the system. For some systems it is the high number of elements that is
relevant for complexity, as in fluid dynamical systems; for others it is the
high number of interactions, as in a small group of swarming animals or
small insect colonies; or it is both, as in the brain and many (if not most)
complex systems. The number of interactions is as important as the number
of elements in the system.
4.2 Disorder and Diversity
Disorder and diversity are related, and the words used to describe them
overlap and are often not clearly defined. ‘Disorder’ usually refers to
randomness, which is to say lack of correlation or structure. Disorder is
therefore just the lack of order. It is worth stating this explicitly since it
follows that any measure of order can be turned into a measure of disorder
and vice versa.
A disordered system is one that is random in the sense of lacking
correlations between parts over space or time or both, at least to some
extent. It is worth remembering that complex systems are never completely
disordered. In complex systems, disorder can exist at the lower level in
terms of the stochasticity in the interactions between the parts, as well as at
the higher level, in terms of the structure which emerges from them and
which is never perfect. Thermal fluctuations are a form of disorder relevant
to the dynamics of complex systems. For example, as noted in Chapter 3,
thermal fluctuations are necessary for most biochemical processes to take
place. The term ‘noise’ or ‘thermal noise’ is used more frequently than
‘disorder’ in this context.
A real or purely mathematical random system would not be described as
‘diverse’. Instead, the term ‘diversity’ is often used to describe
inhomogeneity in element type – that is types of different kinds. Measures
have been designed specifically to address diversity in this sense. Some of
these are discussed at the end of this section.
Interactions can be disordered in time or in their nature. Elements can be
disordered in terms of type. The structure formed by a complex system can
be disordered in its spatial configuration. All these kinds of disorder are
relevant, and all are quantifiable.
Mathematically, disorder is described with the language of probability
(see Section A in the Appendix for a brief introduction to probability
theory). The elements or interactions which are disordered are represented
as a random variable X with probability distribution P over the set of
possible events x (events are elements or interactions). A standard measures
of disorder is the variance. The variance can be used for events that are
numeric, such as the number of edges per node in a network, but not for
types, such as species in a population. The variance of a random variable X,
measures the average deviation from the mean. The equivalent notation in
〉〉
the physics literature is VarX = (X − X )2 . The broader a distribution
of possible event values is the higher, in general, the variance. A second
standard measure of disorder, the Shannon entropy, is a function from
information theory (see Section B in the Appendix for a brief introduction
to information theory). The Shannon entropy of a random variable X with
probability distribution P over events x is defined as
The Shannon entropy measures the amount of uncertainty in the probability
distribution P. In the case of all probabilities being equal, the distribution is
a so-called uniform distribution. In this case all events are equally likely,
and the uncertainty, and hence the Shannon entropy, over events is
maximal. The Shannon entropy is zero when one probability is one and the
others are zero. If, for example, the events x were the possible outcomes of
an election, then H(X) would quantify the difficulty in predicting the actual
outcome.
To illustrate these measures of disorder consider a network, such as the
World Wide Web or a neural network. The disorder relevant to a network is
structural disorder. A network with many nodes and edges between every
pair of nodes is considered a network with no disorder. The origin of a
given network structure is often studied with network-formation models.
For an overview of this and other network-formation models see, for
example, (Newman 2010). One of the first network-formation models is the
so-called Erdös-Renyi random graph model (or just random graph model,
Poisson model, or Bernoulli random graph) (Erdös and Rényi 1960). The
Erdös-Renyi model is parametrised by the number of nodes n, the
maximum number of edges M, and a parameter p which is the probability of
an edge being created between two existing nodes. Initially, the network has
n nodes and no edges. In a subsequent formation process, with probability
p, two nodes are connected by an edge. When p = 0, the resulting network
after many repetitions is a set of nodes without any edges. For p = 1, the
result is a highly connected network. For p somewhere in between 0 and 1,
the formation process yields a network with links between some nodes and
some nodes having more links than others. In this case, the probabilistic
nature of the link formation results in a disordered structure of the network.
Hence, the disorder of the formation process is taken as a proxy for the
disorder of the final network structure. Several properties of the fully
formed network, such as the average path length and the average number of
edges per node, can be expressed as functions of n, M, and p only. These
regularities emerge out of the disorder in the formation process.
The variance can be used to quantify the disorder of the network-
formation process after assigning numeric values to the events ‘edge’ and
‘no edge’ – for example 1 and 0, respectively. The variance of the binary
probability distribution P = {p,1 − p} of the Erdös-Renyi random graph
model is VarX = p(1 − p), which is maximal for p = 1/2. The Shannon
entropy of the network-formation process can be computed without
assigning numerical values to the events. The Shannon entropy of the
binary probability distribution P = {p, 1 − p} of the Erdös-Renyi random
graph model is H(X) = −plog p − (1 − p)log(1 − p), which is also maximal
for p = 1/2. Both measures are zero when p = 0 and, due to symmetry, when
p = 1. If one were to measure the disorder in the final network structure
itself, the variance and the Shannon entropy should be computed from the
probability distribution over the node degrees. The result would be
equivalent to the previous one in the sense that the degree distribution is
trivial for p = 0 (completely disconnected) and p = 1 (fully or nearly fully
connected), in which case both measures yield the value 0. For non-trivial
network structures both measures are non-zero. It was remarked above that
a measure of order can be used as a measure of lack of disorder and vice
versa. Hence, any of the existing measures of network structure, such as
average path length or clustering, can be used to measure disorder by
monitoring their change. This approach to measuring disorder has been
used in the study of Alzheimer’s disease and its effect on neural
connectivity in the human brain (see Bullmore and Sporns (2012) and
references therein).
Temporal disorder in a sequence of events, such as the sequence of daily
share prices on a stock market, is described with the language of stochastic
processes. Disorder in a stochastic process is the lack of correlations
between past and future events. A stochastic process is defined as a
sequence of random variables Xt ordered in time (see Section A of the
Appendix for more details). Disorder is the lack of predictability of future
events when past events are known. To quantify disorder in a sequence
X1X2 ...Xn, the joint probability over two or more of the random variables is
required, written as P(X1X2 ...Xn). This is the probability of the events
occurring together (jointly). When a joint probability of two events is
known, then, in addition to their individual probability, it is known how
likely they are to occur together. An example is the probability of certain
genetic mutations being present and the probability of two mutations being
present in the same genome. The joint Shannon entropy H(X1X2 ...Xn) over
this distribution,
captures the lack of correlations. A measure of average temporal disorder is
the so-called Shannon entropy rate,
The Shannon entropy rate measures the uncertainty in the next event, Xn,
given that all n − 1 previous events X1 ...Xn−1 have been observed. The
lower the entropy rate, the more correlations there are between past and
future events and the more predictable the process is.
A fly’s brain is an example of a complex system where temporal disorder
has been measured experimentally. van Steveninck et al. (1997) recorded
spike trains of a motion-sensitive neuron in the fly’s visual system. From
repeated recordings of neural spike trains, they constructed a probability
distribution P(X1X2 ...Xk) over spike trains of some length k. From this
probability distribution they computed the joint Shannon entropy H(X1X2
...Xk) and the entropy rate . They repeated the
experiments after exposing the fly to the controlled external stimulus of a
visual image and computed the Shannon entropy and entropy rate again.
They interpreted the difference in the entropies between the two
experiments as the reduction in disorder of the neural firing signal when a
stimulus is present.
One speaks of the ‘diversity’ of species in an ecosystem or of diversity
of stocks in an investment portfolio rather than ‘disorder’. In the language
of diversity, the elements, species or stocks, are called ‘types’. The simplest
measure of diversity is the number of types or the logarithm of that number.
A more informative measure takes into account the frequency of each type,
this being the number of individuals of each species in a habitat or the
number of each stock in the portfolio. Treating such frequencies as
probabilities, a random variable X of types can be constructed, and the
Shannon entropy H(X) is used as a measure of type diversity. In ecology,
diversity is measured using the entropy power, 2H(X) (if the entropy is
computed using log base 2 or eH(X) if the entropy is computed using log
base e) (Jost 2006). It behaves similar to the entropy itself but has a useful
interpretation: the entropy power is the number of species in a hypothetical
population in which all species are equally abundant and whose species
distribution has the same Shannon entropy as the actual distribution. If the
types are numeric, such as the size of pups in an elephant seal colony
(Fabiani et al. 2004), diversity can be measured using the variance. Often a
normalised form of variance, the coefficient of variation, is used:
The coefficient of variation is the square root of the variance (also known as
the standard deviation) divided by the mean. Its behaviour is equivalent to
that of the variance. Broader distributions, such as a larger range of pup
sizes in an elephant seal colony, result in a higher coefficient of variation.
However, it allows the comparison of distributions with the same variance
but different means. A distribution with a variance of 10 and a mean of 20
might be considered more diverse than a distribution with a variance of 10
and a mean of 1,000. Their coefficient of variation would reflect this
difference.
Scott Page, in his book ‘Diversity and Complexity’ (2010), distinguishes
between three kinds of diversity: diversity within a type, diversity across
types, and diversity of community composition. All three are measured by
the Shannon entropy. In fact, they differ only in what constitutes an event in
the definition of the random variable. Other measures of diversity are the
so-called ‘distance’ measures and ‘attribute’ measures. Distance measures
of diversity, such as the Weitzman Diversity, take into account not only the
number of types, but also how much they differ from each other (Weitzman
1992) and therefor require a mathematical measure of distance. Attribute-
diversity measures assign attributes to each type and numerically weigh the
importance of each attribute. For example, to compute an attribute diversity
of phenotypes more weight is put on traits with higher relevance for
survival (see Page (2010) for more details).
4.3 Feedback
The interactions in complex systems are iterated so that there is feedback
from previous interactions, in the sense that the parts of the system interact
with their neighbours at later times depending on how they interacted with
them at earlier times. And these interactions take place over a similar time
scale to that of the dynamics of the system as a whole. There is no measure
of feedback as such. Instead, the effects of feedback such as nonlinearity or
structure formation are measured. Hence, the mathematical tools that are
used to measure order and nonlinearity, as described later in the chapter
(Sections 4.5 and 4.6), can also be indicators of feedback.
A common way to study feedback is to construct a mathematical model
with feedback built into it. If the model reproduces the observed dynamics
well, this suggests the presence of feedback in the system that is being
modelled. An example is the dynamics of a population of predator and prey
species such as foxes and rabbits. The growth and decline of these species
can be modelled by the Lotka-Volterra differential equation model. It
describes the change over time in population size of two species, the prey x
and its predator y, using the four parameters A, B, C, and D. A and C are
parameters for the speed of growth of the respective species. B and D
quantify the predation. The change over time in population size, and ,
is given by the two coupled equations
The fact that x and y appear in both equations ensures that there is a
feedback between the size of each population. If B or D are zero, there is no
feedback.
For certain values of the parameters A, B, C, and D the number of
individuals of each species oscillates. When the overabundance of predators
reduces the number of prey to below the level needed to sustain the predator
population but the resulting decline in the number of predators allows the
prey to recover, a cycle of growth and decline results. For such oscillations
to happen the time scale of growth, captured by A and C, needs to be similar
to the time scale of predation, captured by B and D. Oscillations in
predator-prey populations is a classic example of feedback.
A widely used computational tool for studying feedback are so-called
agent-based models. These models are computational simulations of agents
undergoing repeated interactions following simple rules. In such a
simulation a usually large set of agents is equipped with a small set of
actions that each agent is allowed to execute and a small set of (usually
simple) rules defining the interaction between the agents. In any given
round of a simulation an agent and an action, or two agents and an
interaction, are picked at random. If the action (interaction) is allowed, it is
executed. An agent-based simulation usually consists of many thousands of
such rounds. One of the first agent-based models was the sugarscape model,
pioneered by the American epidemiologist Joshua Epstein and
computational, social and political scientist Robert Axtell (Epstein and
Axtell 1996). The sugarscape model is a grid of cells, some of which
contain ‘sugar’; the others contain nothing. Agents ‘move’ on this
landscape of cells and ‘eat’ when they find a cell containing sugar. Even
this very simple setup produces emergent phenomena such as the feedback
effect of the-rich-get-richer, which was described in Sections 2.5 and 2.6 of
Chapter 2.
Agent-based models are frequently used to study feedback in the
coherent dynamics of animal groups (Couzin and Krause 2002). Couzin and
Franks (2003) describe observations of army ants in Soberania National
Park in Panama. Army ants make an excellent study case for collective
phenomena since they are able to form large-scale traffic lanes to transport
food and building material over long distances. They even form bridges out
of ants to avoid ‘traffic congestion’. These collective phenomena are
impossible without the presence of feedback. The authors set up an agent-
based simulation with simple movement and interaction rules for individual
ants. Feedback is built in as an ant’s tendency to avoid collision with other
ants and in its response to local pheromone concentration. The simulations
reproduce the observed lane formations and the minimisation of congestion.
Such a simulation is not to be confused with the measurement of actual
feedback in a real system.
There are other notions of feedback in the literature on complex systems.
The computational notion of feedback is to ‘feed back’ the output of a
computation as input into the same computation. In this way, the outcome
of future computations depends on the outcome of previous computations.
This kind of feedback is particularly important for those who view nature to
be inherently computational (Davies and Gregersen 2014; Lloyd 2006). On
this view, any loop in the computational representation of a natural system
indicates the presence of feedback. Nobel Laureate Paul Nurse made a
similar point when presenting his computational view of the cell (Nurse
2008).
The above tools for analysing feedback have in common that they do not
assign a number to the phenomenon, as is done in the case of disorder or
diversity. Instead, in most practical applications feedback is a tunable
interaction parameter of a model or an observable consequence of the
interactions which are programmed into a model.
4.4 Non-Equilibrium
Complex systems are open to the environment, and they are often driven by
something external. Non-equilibrium physical systems are treated by the
theories of non-equilibrium thermodynamics (De Groot and Mazur 2013)
and stochastic processes (Van Kampen 1992). Stochastic complex systems,
such as chemical reaction systems, are often studied using the statistics of
Markov chains. Consider a system represented by a set of states, S, through
which the system evolves in discrete time steps. Let {Pij} be a matrix of
time-independent probabilities of transitioning from state i to state j, with Σj
Pij = 1,1 for all i ∈ S. Let πi be the probability of being in state i. If there
exists a probability distribution π* such that, for all j,
it is called the invariant distribution. In a stochastic model of a system of
chemical reactions, for example, the chemical composition is represented as
a probability distribution, and chemical reactions are represented as
stochastic transitions from one reactant to another. A system is in chemical
equilibrium if the chemical composition is time-invariant. Reactions are
still taking place in chemical equilibrium, but the depletion of one reactant
is compensated by other transformations such that the overall
concentrations remain largely unchanged. A general framework to model
non-equilibrium stochastic dynamical systems is that of stochastic
differential equations (Ikeda and Watanabe 2014).
4.5 Spontaneous Order and Self-Organisation
Perhaps the most fundamental idea in complexity science is that of order in
a system’s structure or behaviour that arises from the aggregate of a very
large number of disordered and uncoordinated interactions between
elements. Such self-organisation can be quantified by measuring the order
that results – for example, the order in some data about the system.
However, measures of order are not measures of self-organisation as such
since they cannot determine how the order arose. This is because the order
in a string of numbers is the same regardless of its source. Whether the
order is produced spontaneously as a result of uncoordinated interactions in
the system or whether it is the result of external control cannot be inferred
from measuring the order without background knowledge about the system.
For example, the orderly traffic lanes to and from food sources formed by
an ant colony are considered the result of a self-organising process since
there is no mechanism which centrally controls the ants’ behaviour, while
the orderly checkout lines in a supermarket are the result of a centrally
managed control system. A high measure of order, even when self-
organised, is not to be confused with a high level of complexity since order
is but one aspect of complexity. However, the plethora of measures of order
which are labelled as measures of complexity reflects the ubiquity of order
in complex systems and explains the frequent use of order as a proxy for
complexity.
Complex systems can produce order in their environment. It is important
to remember that the order produced by the system is different from the
order in the system itself. For example, the order of hexagonal wax cells
built by honey bees is order produced by the system, while division of
labour in the hive is order in the system. The hexagonal honeycomb
structures are a form of spatial correlation which can be quantified by
correlation measures, some of which are discussed in the following.
A correlation function is a means to measure dependence between
random variables; therefore, it is a statistical measure. The covariance is a
standard measure of correlation. For any two numeric random variables X
and Y, the covariance,
is the difference between the product of the expectations and the
expectation of the product. If the two random variables are uncorrelated,
this difference is zero. From the covariance a dimensionless correlation
measure is derived, the so-called Pearson correlation. It is the most
standard measure of correlation and defined as follows:
where σ is the square root of the variance, known as the standard deviation.
A measure of correlation derived from information theory is the mutual
information. For two random variables X and Y, the mutual information is a
function of the Shannon entropy H (see Section 4.2):
The mutual information measures the difference in uncertainty between
the sum of the individual random variable distributions and the joint
distribution. If there are any correlations between the two variables, the
uncertainty in their joint distribution will be lower than the sum of the
individual distributions. This is a mathematical version of the often repeated
statement that ‘the whole is more than the sum of its parts’. If the whole is
different from the sum of the parts, it means that there are correlations
between the parts. For two completely independent random variables, on
the other hand, H(X) + H(Y ) = H(X,Y) and the mutual information is zero.
An example of covariance as a measure of order is the study of bird
flocking by William Bialek, Andrea Cavagna, and colleagues (2012). They
filmed flocks of starlings in the sky of Rome (containing thousands of
starlings) and extracted the flight paths of the individual birds from these
videos. Each bird’s different flight directions over time were represented as
a random variable, and the random variables of all birds were used to
compute their pairwise covariances.2 This list of covariances was fed into a
computer simulation that modelled the flock of birds as a condensed matter
system, which is defined by the interaction between close-by ‘atoms’ only.
The computer simulation of such a very simple model with pairwise
interactions only and no further parameters, produced a self-organising
system that very closely resembled the self-organising movement originally
observed. A similar analysis was been done on network data of cultured
cortical neurons, corroborating the idea that the brain is self-organising
(Schneidman et al. 2006).
The order in a flock of starlings is a spatial order persistent over time.
Systems in which the focus is more on the temporal aspect of the order are
neurons and their spiking sequences, for example, or the monsoon season
and its patterns. Order in these systems is studied by representing them as
sequences of random variables X1X2 ...Xt with a joint probability
distribution P(X1X2 ...Xt). Such sequences we encountered above in the
study of disorder. Several authors, independently, introduced the mutual
information between parts of a sequence of random variables as a measure
of order in complex systems, under the names of effective measure
complexity (EMC) (Grassberger 1986), predictive information (Ipred)
(Bialek et al. 2001), and excess entropy (E) (Crutchfield and Feldman
2003). Consider the infinite sequence of random variables X−tX−t+1
...X0X1X2 ...Xt, which is also called a stochastic process. The information
theoretic measure Ipred (or EMC or E) of correlation between the two halves
of a stochastic process is defined as the mutual information between the two
halves:
There is, of course, never an infinite time course of data, and the limit t →
∞ is never taken in practice.
Palmer et al. (2015) measured the predictive information in retinal
ganglion cells of salamanders. Ganglion cells are a type of neuron located
near the inner surface of the retina of the eye. In the lab, the salamanders
were exposed, alternatively, to videos of natural scenes and to a video of
random flickering. While a video was showing, the researchers recorded a
salamander’s neural firings. Repeated experiments allowed them to infer the
joint probability distribution P(X−t ...Xt) over the ganglion cell firing rates
and to compute the predictive information Ipred contained in it. They found
that Ipred was highest when a salamander was exposed to naturalistic videos
of underwater scenes. This shows that the order in the natural scenes is
reflected in the order of the neural spike sequences. The authors also think
that it shows the neural system not only responds to a visual stimulus, but
also makes predictions about it.
Quantifying predictability and actually predicting what a system is going
to do are, of course, two different things. In order to make a prediction one
first has to have a model, for example, inferred from a set of measured data.
4.6 Nonlinearity
There are several different phenomena addressed with the same label of
‘nonlinearity’. Each phenomenon requires its own measure. Power laws are
probably the most prominent examples of nonlinearity in complex systems.
But correlations as a form of nonlinearity are equally important, and these
two are not completely separate phenomena.
4.6.1 Nonlinearity as Power Laws
A power law is a relation between two variables, x and y, such that y is a
function of the power of x – for example, y = xμ. Quite a few phenomena in
complex systems, such as the relation between metabolism and body mass
or the number of taxpayers with a certain income and the amount of this
income, follow a power law to some extent. The nonlinear relation between
taxpayer bracket and number of people in this bracket was described in
Chapter 2 in the context of economics. The power law of metabolism for
mammals was first discussed by Max Kleiber (1932) in 1932. It is now well
established that, to a surprising accuracy, the metabolic rate of mammals, R,
is proportional to their body mass, m, to the power of 3/4: R ∝ m3/4 (see
West et al. (1997) and references therein). Because 3/4 is less than 1, a
mammal’s metabolism is more efficient the bigger the mammal; an elephant
requires less energy per unit mass than a mouse. This is a nonlinear effect
since doubling the body size does not double the energy requirements. It is
also another instance of the often repeated, but confused, statement that, in
complex systems, the whole is more than the sum of its parts. The whole is
never more than the sum of its parts when interactions are taken into
account.
The relation between taxpayer bracket and number of people in this
bracket is an instance of a statistical distribution that exhibits a power-law
behaviour. Other examples of statistical distributions with a power-law
behaviour are the number of metropolitan areas relative to their population
size, the number of websites relative to the number of other websites
linking to them, and the number of proteins relative to the number of other
proteins that they interact with (for reviews, see Newman 2005; Sornette
2009).
Statistical distributions with a power-law behaviour are defined in terms
of random variables. Consider a discrete random variable X with positive
events x > 0 and probability distribution P. The distribution P follows a
power law if
for some constant γ > 1 and normalisation constant
, where xmin is the smallest of the x values. A
cumulative distribution with a power-law form is also called a Pareto
distribution (see Section 2.5); a discrete distribution with a power-law form
is also called a Zipf distribution (for a review, see Mitzenmacher 2004). Eq.
4.12 can be written as log P(x) = log c − γ log x, which says that plotting
log P(x) versus log x yields a straight line with slope −γ. Therefore, the
presence of a power law in real-world distributions is often determined by
fitting a straight line to a log-log plot of the data. Although this is common
practice, there are many problems with this method of identifying a power-
law distribution (Clauset et al. 2009).
A power-law distribution has a well-defined mean for γ ≤ 1 over x ∈ [1,
∞) and a well-defined variance for γ ≤ 2. Power-law distributions are
members of the larger family of so-called fat-tailed distributions. These
probability distributions are distinct from the most common distributions,
such as the Gaussian or normal distribution, in that events far away from
the mean have non-negligible probability. Such rare events have obtained
the name ‘black swan’ events; they come as a surprise but have major
consequences (Taleb 2007).
4.6.2 Nonlinearity versus Chaos
Nonlinearity in complex systems is not to be confused with nonlinearity in
dynamical systems. Nonlinear dynamical systems are sets of equations,
often deterministic, describing a trajectory in phase space, either continuous
or discrete in time. Some of these systems exhibit chaos, which is the
mathematical phenomenon of extreme sensitivity of the trajectory on initial
conditions. An example of a discrete dynamical system exhibiting chaos is
the logistic map, already encountered in Chapter 1. The logistic map, xt+1 =
rxt(1 − xt) where t indexes time, is a simple model of population dynamics
of a single species, as opposed to two species, discussed above in the
context of feedback. This map is now a canonical example of chaos.
Actual physical systems studied by dynamical systems theory, such as a
chaotic pendulum, need not have any of the features of complex systems.
Certainly, chaos and complexity are two distinct phenomena. On the other
hand, the time evolution of many complex systems is described by
nonlinear equations. Some climate dynamics, for example, are modelled
using the deterministic Navier-Stokes equations, which are a set of
nonlinear equations describing the motion of fluids. Another example of a
nonlinear equation used to describe many complex systems is the Fisher-
KPP differential equation (Fisher 1937). Originally introduced in the
context of population dynamics, its application ranges from plasma physics
to physiology and ecology.
4.6.3 Nonlinearity as Correlations or Feedback
For some the notion of nonlinearity in complex systems is synonymous
with the presence of correlations (for instance, MacKay 2008). If two
random variables X and Y are independent, their joint probability
distribution P(XY) is equal to the product distribution P(X)P(Y). When this
equality does not hold, then there must be correlations between X and Y.
Defining ‘nonlinearity’ in terms of the presence of correlations is not to
be confused with linear versus nonlinear correlations. In the language of
statistical science, two variables X and Y are linearly correlated if one can
be expressed as a scaled version of the other, X = a + cY, for some constants
a and c. The Pearson correlation coefficient, for example, detects linear
correlations only. The mutual information, on the other hand, detects all
correlations, linear as well as nonlinear.
To others, mainly social scientists, ‘nonlinearity’ means that the causal
links of the system form something more complicated than a single chain. A
system with causal loops, indicating feedback, would count as ‘nonlinear’
in this view (Blalock 1985).
The different definitions of nonlinearity discussed here are all ubiquitous
in complex systems research, so it is not surprising that nonlinearity is often
mentioned as essential to complex systems.
4.7 Robustness
Several phenomena are often grouped together under the umbrella of
‘robustness’. A system might be robust against perturbation in the sense of
maintaining its structure or its function upon perturbation, which some refer
to as ‘stability’. Alternatively, a system might be robust in the sense that it
is able to recover from a perturbation; this is also called ‘resilience’.
Strictly speaking, robustness is the property of a model, an algorithm, or
an experiment that is robust against the change of parameters, of input, or of
assumptions. But usually, in the context of complex systems, robustness
refers to the stability of structure, dynamics or behaviour in the presence of
perturbation. All order and organisation must be robust to some extent to be
worth studying. Several tools are available for studying robustness; the
most frequently used are tools from dynamical systems theory and from the
theory of phase transitions. As with all topics in this chapter, brief
descriptions are given outlining the role of these tools in the study of
complex systems.
4.7.1 Stability Analysis
The system of predator and prey species sharing a habitat, which was
discussed above (see the Lotka-Volterra population model in Section 4.3
and Section 4.6), is an example of a stable dynamical system. After some
time the proportion of the two species becomes either constant or oscillates
regularly, independent of the exact proportion of species in the beginning. A
dynamical system is called ‘stable’ if it reaches the same equilibrium state
under different initial conditions or if it returns to the same equilibrium state
after a small perturbation. Stability analysis is prevalent in physics,
nonlinear dynamics, chemistry, and ecology. A reversible chemical reaction,
for example, might be stable with respect to forced decrease or increase of a
reactant, which means the proportion of reactants and products returns to
the same value as before the perturbation. Other examples of complex
systems which are represented as dynamical systems are food webs with
more than two species (Pimm 1982; Rooney et al. 2006), genetic regulatory
networks (de Jong 2002), and neural brain regions (Friston 2009).
For any given dynamical system described by a state vector x and a set
of (possibly coupled) differential equations dxi/dt = fi(x), a stable point, a
so-called fixed point, is a solution to the equations dxi/dt = 0. Stability
analysis classifies these fixed points into stable and unstable ones (or
possibly stable in one direction and unstable in another). Assuming the
system is at one of its fixed points, the effect of a small perturbation on the
system’s dynamics is found by analysing the Jacobian matrix J, a
linearisation of the system, which is defined as
If the eigenvalues of the Jacobian evaluated at a given fixed point all have
real parts that are negative, then this point is a stable fixed point and the
system returns to the steady state upon perturbation. If any eigenvalue has a
real part that is positive, then the fixed point is unstable and the system will
move away from the fixed point in the direction of the corresponding
eigenvector, usually towards another, stable, fixed point. For an
introduction to fixed-point analysis of dynamical systems, see, for example,
Strogatz (2014). Any stable fixed point is embedded in a so-called basin of
attraction. The size of this basin quantifies the strength of the perturbation
which the system can withstand and, therefore, is a measure of the stability
of the system at the fixed point (Strogatz 2014). Stability analysis is widely
used in ecology (Holling 1973; Scheffer 2010).
Viability theory combines stability analysis of deterministic dynamical
systems theory with control theory (Aubin 1990, 2009). It extends stability
analysis to more general, non-deterministic systems and provides a
mathematical framework for predicting the effect of controlled actions on
the dynamics of such systems, with the aim of regulating them. Viability
theory has been applied to the resilience of social-ecological systems (Béné
and Doyen 2018; Deffuant and Gilbert 2011).
A similar, though mostly qualitative, use of the ideas of stability and
viability is found in the analysis of tipping points in climate and
ecosystems. Tipping points are the points of transition from one stable basin
of attraction to another, instigated by external perturbations (Scheffer 2010).
4.7.2 Critical Slowing Down and Tipping Points
The time it takes for a system to return to a steady state after a perturbation
is a stability indicator complementary to the fixed-point classification and
the size of the attractor basin. The longer it takes the system to recover after
a perturbation the more fragile the system is. An increase in relaxation time
can indicate a critical slowing down and the vicinity to a so-called tipping
point or phase transition. When a system is close to a tipping point, it does
not recover anymore from even very small perturbations and moves to a
different steady state which is possibly very far away from its previous
state. Finding measurable indicators for nearby tipping points has been of
considerable interest, in particular since ecological and climate systems
have begun to be characterised by stability analysis and their fragility is
being recognised more and more (Scheffer 2010; Scheffer et al. 2015).
Mathematically, the vicinity to a tipping point is recognised by the
functional dependency of the recovery time on the perturbation strength. A
system which is close to a tipping point exhibits a recovery time that grows
proportional to perturbation strength to some power. This scaling law,
associated with critical slowing down, is a well-known phenomenon in the
statistical mechanics of phase transitions (see Section 2.1 of Chapter 2 for
more on phase transitions). The standard example of a phase transition in
physics is the magnetisation of a material as a function of temperature. The
magnetisation density m is proportional to the power of the temperature
difference to a critical temperature, m ∝ |T − TC|−a. TC is the critical
temperature, the equivalent to a tipping point, at which the magnetisation
diverges and the system undergoes a phase transition.
Another signature of a nearby tipping point is an increase in fluctuations.
In general, a perturbed system fluctuates around a steady state before
settling back down. The larger the length or time scale on which the
fluctuations are correlated, the closer the system is to a tipping point.
Experimentally, one might expose a system to increasingly strong
perturbations and measure the time it takes the system to come back to its
steady state. Such measurements yield the response of the system as a
random variable S as a function of spatial coordinate x and time t. The
covariance cov(S(x,t),S(x + r,t + τ)) (see Section 4.5) between the random
variable at time t and spatial location x and the same variable at some later
time t + τ and some displaced location x + r is a measure of the temporal
and spatial correlations in time. The equivalent measure in physics is the so-
called autocorrelation function, denoted by C(r,τ), defined, in physics
notation, as
It differs from the covariance by not subtracting the product of the marginal
expectations, S(x,t) 〉 〉
S(x+r,t+τ) (in statistics notation, [S(x,t)]
[S(x+r,t + τ)]), from the expectation of the product (cf. eq. 4.8).
When correlations decay exponentially in time, C is proportional to e−kτ,
where k is an inverse time. After time τ = 1/k correlations have decayed to a
fraction 1/e of the value they had at time t, and τ = 1/k is the so-called
characteristic time scale. Equally, when correlations decay exponentially
with distance, the distance |r| at which they have decayed to a fraction 1/e
of the value at |r| = 0 is the characteristic length scale. Critical slowing
down is accompanied by fluctuations that decay slower than exponentially.
The signature in the auto-correlation function C is a power-law decay either
in time or in space, C ∝ |r|−a or C ∝ τ−a. Theoretically, at the point of a
phase transition the correlation length becomes infinite. At that point the
system has correlations on all scales and no characteristic length nor time
scale anymore. A correlation length which captures nonlinear correlations
has been based on the mutual information (Dunleavy et al. 2015).
An example of a complex system where critical slowing down has been
measured is a population of cyanobacteria under increasing irradiation. The
bacteria require light for photosynthesis, but irradiation levels that are too
high are lethal. For protection against destructively high irradiation levels,
bacteria have evolved a shielding mechanism. Annelies Veraart and
colleagues exposed cell cultures of cyanobacteria to varying intensities of
irradiation and studied the subsequent shielding process (2012). When the
irradiation was relatively weak the bacterial population quickly recovered
after enacting the mutual shielding mechanism by which the bacteria
protect each other. The stronger the radiation, the longer it took the
population to build up the necessary shielding and recover afterwards.
Veraart and her colleagues measured a critical slowing down with a power-
law-like behaviour. Once the light stress reached a certain threshold,
equivalent to a critical point, the population collapsed. The new steady state
that the population had tipped into was that of death. There are many other
complex systems where critical slowing down has been suspected or
observed – for example, in the food web of a lake after introduction of a
predator species (Carpenter et al. 2011), in marine ecosystems in the
Mediterranean after experimental removal of the algal canopy (Benedetti-
Cecchi et al. 2015), and in paleoclimate data around the time of abrupt
climatic shifts (Dakos et al. 2008). For a review of critical slowing down in
ecosystems, see Scheffer et al. (2015). For many more examples of
criticality in complex systems, ranging from geological to financial
systems, see Sornette (2009).
4.7.3 Self-Organised Criticality and Scale
Invariance
Power laws are an example of nonlinearity, as discussed in Section 4.6.
Power-law behaviour is also an example of instability since a power-law
behaviour in the recovery time is the signature of a system being driven
towards a critical point, as discussed. It is, therefore, unexpected that many
complex systems exhibit a power-law behaviour without any visible driving
force and that they are nevertheless relatively stable. It appears that such
systems stay close to a critical point ‘by their own choice’, a phenomenon
called self-organised criticality. When it was discovered in a computer
model of avalanches (Bak et al. (1988); see Section 2.1), it sparked a whole
surge of studies into the mechanism behind self-organised criticality. This
surge was fueled by experimental observations of power-law-like behaviour
in a range of different systems, such as the Earth’s mantle and the
magnitude and timing of earthquakes and their afterquakes, or the brain and
the timing of neurons (Bullmore and Sporns 2009; Sornette 2002; Zöller et
al. 2001). In these systems, the relevant observable, magnitude or timing,
was measured as a histogram of frequencies of events. The probability
distribution P(x) of events x, constructed from the data, decays
approximately as a power law, P(x) = cx−γ. As remarked above, the true
functional form of these decays is still debated; it is rarely more than an
approximate power law (Clauset et al. 2009). A power law implies so-called
scale invariance, since ratios are invariant to scaling of the argument:
P(cx1)/P(cx2) = P(x1)/P(x2). Scale invariance has been observed in many
natural as well as social complex systems (Sornette 2009), including scale
invariance of the statistics of population sizes of cities (Bettencourt 2013).
While a power law in the auto-correlation function indicates instability
and the vicinity of a critical point, a power law in a statistical distribution
may indicate self-organised criticality which is associated with stability.
Three decades after the discovery of self-organised criticality, there still is
no known mechanism for it. The seeming contradiction between the
robustness of a complex system, one of its emerging features, and the
inherent instability of systems close to a tipping point remains unresolved.
For a review of self-organised criticality, see Pruessner (2012) and Watkins
et al. (2016).
4.7.4 Robustness of Complex Networks
Network structures are ubiquitous in the interactions within a complex
system. It is therefore not surprising that complex networks have grown
into their own subfield of complex systems research. Many examples of
networks have been mentioned in this book, from protein-protein
interactions and neural networks to financial networks and online social
networks. A network is a collection of nodes connected via edges. The
degree of a node is the number of edges connected to it. The nature of
nodes and edges differs for each system. In protein-protein networks the
nodes are proteins; two nodes are connected by an edge if they interact,
either biochemically or through electrostatic forces. A path is a sequence of
nodes such that every two consecutive nodes in the sequence are connected
by an edge. The path length is the number of edges traversed along the
sequence of a path. The average shortest path is the sum of all shortest path
lengths divided by their number. The phrase ‘six degrees of separation’
refers to the average path length between nodes in social networks. This
goes back to a now famous experiment performed by Stanley Milgram and
his team in the 1960s (Milgram 1967). Milgram gave letters to participants
randomly chosen from the population of the United States. The letters were
addressed to a person unknown to them, and they were tasked with handing
their letter to a person they knew by first name and who they believed
would be more likely to know the recipient. This led to a chain of passings-
on for each letter. Surprisingly, letters reached the addressee, on average,
after only five intermediaries. The stability of average path length is one
proxy for the robustness of a network. When edges or nodes are removed
from the network and the average path length stays more or less the same,
the network is considered robust (in this respect). Reka Albert, Hawoong
Jeong and Albert-Lásló Barabási (2000) found that the Internet and the
World Wide Web are very robust in precisely this way. The shortest path is
hardly affected upon the random removal of nodes. Albert and her
colleagues studied the structure of the World Wide Web and the Internet by
taking real-world data and artificially removing nodes in a computer
simulation. Plotting the shortest path against the fraction of nodes removed
from the network revealed that the path length initially stayed
approximately constant. Only once a large fraction of the nodes had been
removed did the length suddenly and dramatically increase. This sudden
increase is a form of phase transition between a well-connected phase and a
disconnected phase. It is seen already in the simplest model of networks,
the Erdös-Renyi random graph model discussed above (see Newman 2010).
Other real-world networks exhibiting this structural form of robustness are
protein networks (Jeong et al. 2001), food webs (Dunne et al. 2002), and
social networks (Newman et al. 2002). Robustness is always with respect to
a feature or function. Robustness with respect to one feature might not
imply robustness with respect to another. The Internet, for example, is
robust against random removal of nodes (servers), but it is considerably less
robust to targeted removal of the highest-degree nodes.
4.8 Nested Structure and Modularity
Nested structure and modularity are two distinct phenomena, but they may
be related. ‘Nested structure’ refers to structure on multiple scales.
‘Modularity’ is a functional division of labour, or specialisation of function
among parts, or a structural modularity and frequently all of these together.
Structural modularity is a property much discussed especially in the
context of networks, where it is referred to as ‘clustering’. A cluster in a
network is a collection of nodes that have many edges between one another
compared to only few edges to nodes in the rest of the network. A simple
example is the network of online social connections such as the network of
‘friends’ on Facebook. This network of social connections tends to be
highly clustered since two ‘friends’ of any given user are more likely to also
be ‘friends’ than to be unrelated.
Finding clusters in networks has received considerable attention, and
many so-called clustering algorithms have been proposed. For an
introduction to clustering algorithms, see, for example, Newman (2010). All
clustering algorithms follow a similar principle. Given a network, they
initially group the nodes into arbitrary communities, and, by some measure
unique to each technique, they quantify the linking strength within each
community and that in between communities. Information theoretic
distance is one such measure (Rosvall and Bergstrom 2008). The algorithms
then optimise the communities by moving nodes between them until the
linking strength within each community is maximised and the linking
strength in between communities is minimised. There is usually no unique
solution to this optimisation problem, and the identified clusters might
differ from algorithm to algorithm. The presence of clusters alone is not
sufficient for modularity since the network could consist of one gigantic
cluster, with every node being connected to most other nodes, and have no
modularity at all.
Once a community structure of a network has been identified, the extent
to which it is modular can then be quantified. One of the first measures
designed to quantify structural modularity is the modularity measure by
Mark Newman and Michelle Girvan (2004). It assumes that a community
structure of a given network has been identified and that k clusters of nodes
have been found. From these k clusters, a k × k matrix e is constructed in
which the entries eij are the fraction of edges that link nodes in cluster i to
nodes in cluster j. The matrix entries can be interpreted as the joint
probabilities Pr(i, j) for the event of an edge to be attached to a node in
cluster i and the joint event of this edge to end on a node in cluster j. If
these two events are independent, the joint probability distribution is equal
to its product distribution, Pr(i, j) = Pr(i) · Pr(j). If, on the other hand, Pr(i,
j) ≠ = Pr(i) · Pr(j), then the probability Pr(i, j) is dependent on whether i and
j are the same cluster (i = j) or not. With such a dependence present, there is
modularity in the network. This condition of a joint probability distribution
being a non-product distribution was a condition for ‘nonlinearity as
correlations’ (Section 4.6.3).
Newman and Girvan use this dependency condition to define modularity
Q as any deviation of the joint probability distribution Pr(i,i) of edges
connecting nodes within the same cluster from the product distribution Pr(i)
·Pr(i). In this sense, modularity is a form of nonlinearity as correlations. In
the above matrix notation, the probability Pr(i) = Σj eij. This can be
understood as the so-called marginal probability of picking any edge in the
network and for that edge to start in cluster i. Modularity is then defined as:
This measure of modularity is also taken as an optimisation function for
community detection algorithms, but limitations to its effectiveness have
been pointed out (Brandes et al. 2007; Fortunato and Barthelemy 2007).
Many natural systems exhibit structure that is repeated again on a
smaller scale; the structure is nested within itself. A cauliflower exhibits
this particular form of spatial scale invariance in the structure of the florets
consisting of smaller florets and so forth. Benoît Mandelbrot discovered the
mathematics of such nested structures, for which he coined the term fractal.
Fractals are mathematical objects with a perfect scale invariance, a
repetition of structure at an infinite number of scales (Falconer 2004).
Mandelbrot’s now famous book, The Fractal Geometry of Nature (1983),
revealed the ubiquitous presence of fractal structure in natural systems, both
living and nonliving. Fractals have the mathematical property of a non-
integer dimension, and therefore fractal dimension is sometimes used as an
indicator of nested structure (e.g., in ecology; Sugihara and May 1990). For
example, a circle has dimension 2, a sphere has dimension 3, and the
dimension of a cauliflower is estimated at 2.8 (Kim 2004).
Another indicator for multiple scales is the power-law decay in a
correlation function (see Section 4.7). For example, as mentioned in Section
2.6 of Chapter 2, the number of websites in the visible World Wide Web as
a function of their degree approximately follows a power law with an
exponent γ, which, in 1999, was estimated at 2.1 (Barabási and Albert
1999). This power-law decay is due to clusters of websites being nested
within bigger clusters of websites. The World Wide Web has tens of billions
of web pages, but only a few dozen domains own most of the links.3 These
central domains are linked to each other, as well as to web pages within
their own domain, and they also connect to large clusters of less-well-
connected domains. Each of these clusters has, again, a few highly
connected domains. This structure of clusters of sites with a few highly
linked domains repeats at ever smaller scales. This self-similar nesting of
clusters is much studied in complex networks (Newman 2003; Ravasz and
Barabási 2003). Methods based on statistical inference for identifying
nested clusters have also been developed (Clauset et al. 2008).
The presence of scale invariance in the degree distribution of a network
can be reproduced by a model of network growth first considered by Derek
Price (1976). Starting with a small network, new nodes are added and
connected by an edge to an existing node with a probability proportional to
the existing node’s degree. Hence, any new edge will affect the probability
of future edges being added. Connecting a large number of nodes following
this rule results in a network where a few nodes have a very high number of
edges, and most nodes have very few. The algorithm is called the
preferential attachment algorithm. It is a variant of a random graph model
(see Section 4.2), and it describes the rich-get-richer effect seen in
economics (see Section 2.5 of Chapter 2). It clearly has feedback built into
it. The initial degree distribution might be uniformly random, but, after
many iterations, it gets locked into a very skewed distribution due to the
feedback of previously formed edges on future edge formation. The
preferential attachment mechanism illustrates why power laws have been,
and still are, a central theme in many studies of complex systems. Power-
law-like behaviour, already discussed repeatedly in this chapter, can serve
as an indicator for several of the features of complex systems identified in
this book: nonlinearity, (lack of) robustness, nested structure, and feedback.
This also suggests that these phenomena are not isolated from each other.
4.9 History and Memory
The various measures of complexity measure different features of complex
systems, all of which arise because of their histories. Hence, other measures
can be used as proxies for history. For example, a network may have a
definite growth rate so that the size of the network can be used as a measure
of its age. Another way to measure the history of a complex system is to
measure the structure it has left behind, because the more of it there is, the
longer the history required for it to spontaneously arise as a result of the
complex system’s internal dynamics and interaction with the environment.
However, in some cases, the structure in the world is built very quickly and
deliberately rather than arising spontaneously, like a beaver’s dam or a
ploughed field. Background knowledge is needed to know how to relate
such structure to history. There are no direct measures of history used in
practice, but the logical depth discussed in the next section was introduced
to capture the idea that complex systems require a long history to develop.
Any measure of correlations in time, including the statistical complexity
discussed below, can be considered a measure of memory.
4.10 Computational Measures
Many of the growing number of measures of complexity are based on
computational concepts such as algorithmic complexity and
compressibility.4 The previous section showed that complexity measures
capture features of complexity but not complexity as such. This section
discusses measures that consider complex systems to be computational
devices with memory and computational power. All of these measures are
reminiscent of thought experiments in that they are not implementable in
practice or even in principle. Although these measures are now decades old
(and none measure complexity as such) they are included here because they
have had a considerable influence on thinking about complexity. We
explain what feature of complexity each measures.
4.10.1 Thermodynamic Depth
Thermodynamic depth was introduced by physicists Seth Lloyd and Heinz
Pagels (1988). Lloyd and Pagels started out with the intuition that a
complex system is neither perfectly ordered nor perfectly random and that a
complex system plus a copy of it is not much more complex than one
system alone. To specify the order of a complex system they consider the
physical state of the system at time tn, calling it sn. In any stochastic setting,
a given state can be preceded by more than one state. In other words, the set
of states a system was in at times t1 to tn−1, a trajectory of length n − 1, is
not unique. Assigning a probability to any such trajectory which leads to
state sn, Pr(s1,s2,...,sn−1|sn), the thermodynamic depth of state sn is defined
as −k lnPr(s1,s2,...,sn−1|sn) averaged over all possible trajectories
s1,s2,...,sn−1,
where k is the Boltzmann constant from statistical mechanics. In this view,
the complexity of a system is given by the thermodynamic depth of its state.
The intuition that the thermodynamic depth is intended to capture is that
systems with many possible and long histories are more complex than
systems which have short, and thus necessarily fewer possible, histories.
What this definition leaves open, and arguably subjective, is how to find the
possible histories, their lengths and what probabilities to assign to them
(Ladyman et al. 2013). Thus, practically, the measure is not implementable.
The rate of increase of thermodynamic depth when considering histories
further and further back in time is mathematically an entropy rate, which is
a measure of disorder (see Section 4.2). Thus, while the intention was for
thermodynamic depth to be a measure of history, it is in fact a measure of
disorder. This was pointed out in Crutchfield and Shalizi (1999).
4.10.2 Statistical Complexity and True Measure
Complexity
The quantitative theory of self-generated complexity, introduced by
physicist Peter Grassberger (1986), and computational mechanics,
introduced by physicists James Crutchfield and Karl Young (Crutchfield
1994; Crutchfield and Young 1989), are similar frameworks that go beyond
providing a measure to inferring a computational representation for a
complex system. The former comes with a measure called true measure
complexity, the latter with a measure called statistical complexity. Since
computational mechanics has been developed in more statistical and
practical detail, we focus on it here (see Shalizi and Crutchfield 2001).
The assumption of computational mechanics is that a complex system is
an information-storing and -processing entity. Hence, any structured
behaviour it exhibits is the result of a computation. The starting point of the
inference method is a representation of the system’s behaviour as a string
such as, for example, a time sequence of measurements of its location.5 The
symbols in this measurement sequence generally form a discrete and finite
set (for background, see Section A in the Appendix). Once a string of
measurement data has been obtained, the regularities are extracted using an
algorithm which is briefly explained below, and a computational
representation is inferred which reproduces the statistical regularities of the
string. Computational representations can, in principle, be anything from
the Chomsky hierarchy of computational devices (Hopcroft et al. 2001), but
in concrete examples they usually are finite-state automata. The size of this
automaton is the basis for the statistical complexity measure.
The algorithm for inferring the computational representation of a string
assumes that a stationary stochastic process has generated the
string in question (for a definition of stationary stochastic process, see
Section A of the Appendix). As a next step, statistically equivalent strings
are grouped together. Two strings, and , are statistically equivalent if
they have the same conditional probability distribution over the subsequent
symbol a ∈ :
The two strings do not have to be of the same length. The equivalence class
of a substring is denoted by ɛ( ), and it contains all strings statistically
equivalent to string , including itself. These classes are called ‘causal
states’, a somewhat unfortunate name since no causality is implied in any
strict sense. Due to the stationarity of the process, the transition
probabilities between the causal states are stationary and form a stochastic
matrix. Hence, the computational representation obtained by this algorithm
is a stochastic finite state automaton or, equivalently, a hidden Markov
model (Hopcroft et al. 2001; Paz 1971) and is called ɛ-machine.6
The stationary probability distribution P of the ɛ-machine’s causal states
s∈ , which is the left eigenvector of its stochastic transition matrix with
eigenvalue 1, is used to define the statistical complexity, Cμ, of a process:
where is the set of causal states. Cμ is the Shannon entropy of the
stationary probability distribution. This reflects the computational
viewpoint of the authors since, technically, the Shannon entropy is the
minimum number of bits required to encode the set with probability
distribution P. Thus, the statistical complexity is a measure of the minimum
amount of memory required to optimally encode the set of behaviours of
the complex system. It is worth noting that, for a given string, the statistical
complexity is lower bounded by the excess entropy/predictive information
(see eq. 4.11 above), Cμ ≥ Ipred (eq. 4.11) (Crutchfield and Feldman 2003;
Crutchfield et al. 2009; Wiesner et al. 2012). This mathematical fact agrees
with the intuition that a system must store at least as much information as
the structure it produces. The statistical complexity has been computed for
the logistic map (see Chapter 1) (Crutchfield and Young 1989), for protein
configurations (Kelly et al. 2012; Li et al. 2008), atmospheric turbulence
(Palmer et al. 2000), and for self-organisation in cellular automata (Shalizi
et al. 2004).
Crutchfield (1994, p. 24) writes that “an ideal random process has zero
statistical complexity. At the other end of the spectrum, simple periodic
processes have low statistical complexity. Complex processes arise between
these extremes and are an amalgam of predictable and stochastic
mechanisms.” This statement, though intuitive, is obscuring the fact that the
statistical complexity increases monotonically with the order of the string.
For a proof consider the following. For a given number of causal states the
statistical complexity has a unique maximum at uniform probability
distribution over the states. This is achieved by a perfectly periodic
sequence with period equal to the number of states. When deviations occur,
the probability distribution will, in general, not be uniform anymore, and
the Shannon entropy and with it the statistical complexity will decrease. On
the other hand, increasing the period of the sequence requires an increased
number of causal states and, thus, implies a higher statistical complexity.
Hence, the statistical complexity scores higher for highly ordered strings
than for strings with less order or with random bits inserted. The statistical
complexity is a measure of order produced by the system, as well as a
measure of memory of the system itself. The strength of the framework of
computational mechanics lies in detecting order in the presence of disorder.
4.10.3 Effective Complexity
Effective complexity was introduced by physicists Murray Gell-Mann and
Seth Lloyd (1996). Gell-Mann and Lloyd’s starting point is common to
many measures of complexity of that time: the measure should capture the
property of a complex system of being neither completely ordered nor
completely random. They assume the complex system can be represented as
a string of bits; call it s. This string is some form of unique description of
the system or of its behaviour or the order it produced. The algorithmic
complexity (for a definition, see Section C in the Appendix) of this string of
bits is a measure of its randomness or lack of compressibility. The more
regularities a string has, the lower is its algorithmic complexity. Hence,
Gell-Mann and Lloyd consider the algorithmic complexity not of the string
itself, but of the ensemble (a term taken from statistical mechanics) of
strings with the same regularities as the string in question. Let E be this
ensemble of strings with the same regularities. The effective complexity of
the string (and thus the system which it represents) is defined as the
algorithmic complexity of the ensemble E in which it is embedded as a
typical member. (‘Typical’ is a technical term here, but it captures exactly
what we intuitively think ‘typical’ should mean.) Ensemble members s ∈ E
are called typical if −logPr(s) ≈ K (E), where K (E) is the algorithmic
complexity of E (see Section C of the Appendix).7 Assigning a probability
to each string in a set is less arbitrary than it sounds. It has been shown that
the probability Pr(s) of a string s is related to its algorithmic complexity as
−logPr(s) ≈ K (s|E) where K (s|E) is the algorithmic complexity of
the string s given a description of the set E. The effective complexity ɛ(s) of
a string s is defined as the algorithmic complexity of the ensemble E of
which it is a typical member,
For example, the ensemble of a string which is perfectly random is the set
of all strings of the same length. This set allows for a very short description,
by giving the length of the strings only. This trick of embedding the string
in a set of similar strings exactly achieves what Gell-Mann and Lloyd set
out to do. A string with many regularities over many different length scales,
which is how we think of a complex system, will be assigned a high
effective complexity. Random systems, in their structure or behaviour, will
be assigned very low effective complexity. According to Gell-Mann and
Lloyd (1996), the effective complexity can be high only in a region
intermediate between total order and complete disorder. However, replacing
some of the regular bits in a string by random bits decreases its regularities
and hence its effective complexity. Just like the true measure complexity
and the statistical complexity, the effective complexity increases
monotonically with the amount of order present. This places the effective
complexity among the measures of order. Gell-Mann and Lloyd note that
this measure is subjective, since what to count as a regular feature is the
observer’s decision. The instruction “find the ensemble (of a rain forest, for
example) and determine its typical members” leaves too many things
unspecified for this measure to be practicable (McAllister 2003).
4.10.4 Logical Depth
Computer scientist Charles Bennett introduced logical depth to measure a
system’s history (Bennett 1991). Bennett argues that complex objects are
those whose most plausible explanations involve long causal processes.
This idea goes back to Herbert Simon’s influential paper, ‘The Architecture
of Complexity’ (1962). To develop a mathematical definition of causal
histories of complex systems, Bennett replaces the system to be measured
by a description of the system, given as a string of bits. This procedure
should be very familiar by now. He equates the causal history of the system
with the algorithmic complexity of the string (the length of the shortest
program which outputs the string; see Section C of the Appendix). The
shorter the program which outputs a system’s description, the more
plausible it is as its causal history. A program consisting of the string itself
and the ‘print’ command has high algorithmic complexity, and it offers no
explanation whatsoever. It is equivalent to saying ‘It just happened’ and so
is effectively the null-hypothesis. A program with instructions for
computing the string from some initial conditions, on the other hand, must
contain some description of its history and thus is a more explanatory
hypothesis.
In addition to considering a program’s length as a measure of causal
history, Bennett also takes the program’s running time into account. A
program which runs for a long time before outputting a result signifies that
the string has a complicated order that needs unravelling. The definition of
logical depth is then as follows. Let x be a finite string, and K (x) its
algorithmic complexity. The logical depth of x at significance level s is
defined as the least time T(p) required for program p to compute x and then
halt where the length of program p, l(p), cannot differ from KU (x) by more
than s bits,
Logical depth combines the features of order and history into a single
measure. Consider the structure of a protein, for example. One possible
program prints the electron and nuclear densities verbatim, with discretised
positional information, which is a very long program running very fast.
Another program computes the structure ab initio by running quantum
chemical calculations. This would be a much shorter program but running
for a long time. The latter captures the protein’s order and history. The real
causal history of a protein is, of course, very long, starting with the
beginning of life on Earth, or even with the beginning of the universe. The
logical depth captures our intuition that complex systems have a long
history. A practical problem is that the time point when the history of a
system starts is not well defined. Another aspect which makes it impractical
to use is that the algorithmic complexity is uncomputable in principle,
although approximations exist such as the Lempel-Ziv algorithm (Ziv and
Lempel 1977). The next chapter returns to the question of ‘What is a
complex system?’
1Some scientific fields use the reverse order, {P }.
ji
2They used the convention from statistical mechanics in which the uncorrelated average product
is not subtracted. Thus, their covariance is the statistical mechanical correlation function E[XY].
3A domain is the name you need to buy or register, such as google.com. Any web page with a url
within this domain name, such as www.google.com/maps is part of this domain.
4Lloyd (2001) produced a ‘non-exhaustive list’ of over forty, and many more measures have been
defined since then.
5Of course, measurement is crucial to computational mechanics, and it raises many practical
questions left aside here.
6The inference algorithm is available in various languages, see, for example, Computational
Mechanics Group (2015); Kelly et al. (2012); Shalizi and Klinkner (2003).
7The idea of replacing entropy (the average of −logPr(s) is an entropy) by algorithmic complexity
goes back to Wojciech Zurek (1989).
Chapter 5
What Is a Complex System?
Is there a single natural phenomenon of complexity found in a wide variety
of living and nonliving systems and which can be the subject of a single
scientific theory? Is there such a thing as ‘complexity science’ rather than
merely branches of different sciences, each of which have to deal with their
own examples of complex systems? This chapter synthesises an account of
how to think about complexity and complex systems from the examples and
analysis of the preceding chapters. Roughly speaking, our answers to these
questions are no to the first and yes to the second. There is no single
phenomenon of complexity, but there are a variety of features of complex
systems that manifest themselves in different ways in different contexts.
Hence, complexity science is not a single scientific theory but a collection
of models and theories that can be used to study the different features in
common ways across different kinds of systems. The following sections
consider different views about complex systems, and the penultimate
section argues for our view. The final section of this chapter reflects on the
broader implications of what has been learned.
5.1 Nihilism about Complex Systems
The most negative view of complexity is simply that there is no such thing.
Similarly, it could be argued that ‘complex system’ is a vague and
ambiguous term that covers a variety of things and that complexity science
is just a collection of techniques and methods that does not have a domain
of its own. In this way of thinking, which we can call ‘nihilism’, there are
very different ways that different sciences are combined to study
complicated systems, and complex systems are just complicated things we
study with interdisciplinary science using computers. Anything over and
above that is, at best, a convenient label for an emerging synthesis (in the
sense discussed in Chapter 1) and, at worst, hype for the purposes of grant
applications, journal publications and book sales. On this kind of view, it
seems advisable to stop using the terms ‘complexity’ and ‘complex system’
as if they were well-defined scientific concepts.
Consider nihilism in the light of the discussion in Chapter 2. For
example, take the human brain. It is ultimately subject to the laws of
physics, chemistry and biochemistry. All the basic processes that occur in it
can be described in the language of these existing scientific disciplines. It
might be argued that the necessity to describe the brain as a complex
system, and to use the new techniques of complexity science, is simply due
to the fact that the physical and chemical mechanisms relevant for its
behaviour are so numerous and complicated that it is impossible to describe
them all in practice. Once one starts to care about details, underlying
mechanisms and scientific explanations of them, the brain necessarily
enters the realm of physics, chemistry and biochemistry. So described, the
brain is the subject of each of the traditional sciences, and it might be
supposed it could be completely described in the language of existing
physics and chemistry. Unless there is a fundamental new set of forces and
particles associated with the brain, for which there is no evidence, the brain
is a physical and biochemical system. Theories of the brain ultimately rely
on basic science and fit into the hierarchical organisation of nature above
physics, chemistry and biology. Similarly, for the rest of the examples so it
might be argued, there is no single concept of complex system; there are
just systems that are physical, chemical, biological, economic, or whatever.
Of course, nihilists must admit that the complex systems perspective has
allowed problems to be discovered and discussed in an unconventional,
interdisciplinary way. For example, scientists now study the immune system
as a network of nodes and links. However, they do not necessarily regard
this as a new science, because this perspective is embedded into the existing
frameworks of anatomy and physiology. In this view, there are two possible
paths that a new problem and its solution can take in the scientific
disciplinary landscape: either it can be completely swallowed by the
existing field as if it had always been part of it, or it can become a field in
its own right, such as neuroscience, which uses ideas from network theory.
Either way, the problems and ansatz to solutions only remain part of the
field of complex systems until the system has been fully understood.
According to nihilism, the field of complexity science is just a feeder of
new science to other fields but not a science in its own right.
This view does not directly contradict any of the truisms of Chapter 1,
but it does not take account of the fact that there are kinds of invariance and
universal behaviour studied by complexity science that are found in many
very different kinds of systems. Furthermore, as shown in the last chapter,
there are novel kinds of invariance and forms of universal behaviour that
are found when complex systems are modelled as networks and
information-processing systems. Hence, nihilism about complex systems
should be rejected.
5.2 Pragmatism about Complex Systems
A different view agrees with nihilism that there is no essence of complexity
or complex systems but is nonetheless much more positive about
complexity science. ‘Pragmatism’ says that there is enough of a
resemblance between the collection of methodologies, mathematical and
computational, which are applicable to a whole variety of systems, from
proteins to the brain to forests and cities, in order for the ideas of
complexity and complex systems to be scientifically useful. In this
approach, the term ‘complex system’ refers to a collection of systems in
various disciplines that are all amenable to related mathematical and
computational techniques, such as network theory or agent-based
modelling, but which otherwise have nothing fundamental in common.
According to this approach, the wide applicability of complexity science is
not necessarily indicative of any unity to the idea of a complex system. The
techniques are useful to scientists, so they are worth teaching to students,
and this is sufficient to justify the existence of the many research centres for
complexity science throughout the world. Pragmatism says complexity
science is a range of models, theories and techniques that are applicable
very widely. Strevens (2016) advocates pragmatism about complexity
science in this sense.
Pragmatism is probably a common view among complexity scientists
who have no interest in worrying about what complexity is or what
complex systems are.1 However, there are many who go further and argue
for particular definitions of complexity and complex systems or at least who
impose necessary conditions that rule out some cases. Some of these
conceptions of complexity are considered below (they all originate with
scientists working in the field). The pragmatist about complex systems need
not deny any of these views but may be agnostic about them.
5.3 Realism about Complex Systems
Realism about complex systems is the idea that ‘complex systems’ form
what philosophers call a ‘natural kind’. Most scientists probably think that
the elements of the periodic table, or the fundamental particles that make
them up, are the ultimate natural kinds. In science, it is crucial to find the
right way of dividing things up into kinds. For example, whales should not
be put in the same category as fish just because they live in the sea and have
fins, and jade is not in fact a natural kind of mineral (because there are two
different minerals, jadeite and nephrite, that have both been given the same
name). Every science has its own taxonomy. There are white dwarfs and red
giants in astrophysics, noble gases and halogens in chemistry, and
organisms and classes of them in biology. Often the way things are
classified is refined over time, and arriving at an exact definition may be
difficult or impossible. For example, the definition of an acid has undergone
various revisions, and there is no consensus about what exactly life is and
whether viruses are alive. This implies that if the concept ‘complex system’
is vague, that does not make it any less scientifically valid than concepts
such as that of life. The realist can argue that just as living systems form a
natural kind, so too do complex systems, but only if they are able to say
something, vague or not, about what complex systems are.
It is important to realism about science in general that there are natural
properties that are independent of our conventions and interests. For
example, the mass of objects or their electric charge is what it is, however
we decide to quantify them with a choice of units. One version of the realist
view of complex systems, popular with many complexity scientists, is that
there is such a natural property of complexity that can be quantified and
measured. Then complex systems are derivatively just those that have a lot
of complexity.
As discussed at the end of Chapter 1, scientists who hold the realist view
are not at all a uniform group. Some understand complexity narrowly so
that only biological systems count as complex. Others – for example,
Murray Gell-Mann (1994) – understand it more broadly so that physical and
chemical systems that are not associated with life can also be complex. As a
result, among the realist conceptions of complexity that are proposed in the
literature, some are couched in generic scientific language, some involve
only physical quantities, some use computational/information-theoretic
concepts, and some require biological ideas. Below, a representative sample
of views is outlined.
5.3.1 Generic Conceptions of Complexity
Generic conceptions of complexity are those that can be stated using
generic scientific or mathematical language (often that of dynamical
systems theory). In particular, they do not make reference to physical
quantities, computational or information-theoretic ideas, or biological
concepts.
Complexity as Emergence
As defined in Chapters 1 and 2, the most basic kind of emergence is the
existence of laws and properties at the level of the whole system that do not
exist at the level of the constituent parts. As stated at the end of Chapter 3,
the simplest realist view is that complex systems are all and only those that
exhibit emergence. This view reflects the fact that everyone in the
complexity science community takes emergence to be central and has done
so ever since the first meetings in Santa Fe, as described in Chapter 1. This
view is clear and can be made exact, because it is possible to describe
mathematically how a system with very many degrees of freedom can have
a much simpler lower-dimensional dynamics defined on effective states, as
in statistical mechanics.
Emergence can be purely epistemological, meaning the emergent entities
do not really exist but are just convenient ways to keep track of what is
happening in the system. For example, the Game of Life shows a merely
epistemological kind of emergence because ‘eaters’ and ‘gliders’ do not
really eat or glide respectively; they are just a kind of useful fiction for
keeping track of the evolution of the system. On the other hand, the
emergence from the physical systems in the universe to the immensely
intricate and structured system of life on Earth, including the human brain
and the complexity of human culture and social life, is ontological
emergence (unless there are not really animals and people and so on).2
Certainly emergence in all epistemological senses is necessary for a
complex system. If a system does not exhibit higher-level order of some
kind that arises spontaneously from the interactions of its parts, then it is
not complex. However, as discussed in Section 2.1, even an isolated gas at
thermal equilibrium has emergent properties of pressure and temperature
that an individual molecule does not. Such a gas lacks all the interesting
features of complex systems described in detail in Chapters 3 and 4 other
than this most minimal kind of emergence, which is exhibited by everything
in the sciences except perhaps the entities of the ultimate fundamental
physics if it exists.
Since there are different kinds of emergence, it may be argued that some
specific kind of emergence is required of a complex system. This view may
then coincide with one of the other views discussed below, depending on
what kind of emergence is taken to be necessary.
Complexity between Order and Chaos
Many people say that a system is complex if it lies between order and chaos
(Crutchfield 2012). As argued in Chapter 3, the notions of order and
disorder are crucial to the understanding of complex systems, but they are
not sufficient for a definition. The idea of complexity as some kind of
middle ground between order and disorder is also misleading. A crystal
with many impurities, which is a very ordered arrangement of atoms or
molecules with some atoms slightly misplaced or replaced by different
atoms, is not thereby more complex than a crystal with fewer impurities or
more complex that a crystal with more impurities. Accounts of complexity
based on ‘measures of complexity’ that measure order in one way or
another confuse order with the way that it is produced, as argued in Chapter
4. Furthermore, as argued in Chapter 3, ideas of order and disorder are
always relative to some way of representing the system or gathering data
about it, and what is ordered at one scale may be disordered at another and
vice versa. The analysis of the previous chapters shows that complexity is
too diverse to be captured in this way.
5.3.2 Physical Conceptions of Complexity
Physical conceptions of complexity are those that use only the language of
physics to characterise complexity. In particular, they do not make reference
to computational or information-theoretic ideas or to biological concepts.
Note that biological systems can count as complex systems (and could even
be the only ones there are) according to physical conceptions of complexity.
The important thing is that the following conceptions of complexity do not
define it in terms of biological concepts, even if only living things satisfy
such a definition.
Complexity as Non-Equilibrium Thermodynamics
Various authors characterise complexity in terms of thermodynamic
properties such as entropy, energy and being out of equilibrium (see
Chapter 3). For example, astrobiologist Charles Lineweaver (2013)
explores the idea that the creation of complexity is driven by free energy
and argues that, as the latter decreases with the increase of entropy, so too
will complexity decrease as we head towards the heat death of the universe.
The association of complexity with free energy is illustrated by the fact that
the degree of order of a hurricane depends on pressure, temperature and
humidity gradients that correspond to the gradient of free energy (recall also
the discussion of Section 3.4. The formation and maintenance of biological
structure also seems to require a free energy gradient. On the other hand,
the astrophysicist Eric Chaisson (2013) proposes energy flow per unit time
per unit mass as a measure of complexity. He applies this idea to both
nonliving and living systems to argue that in both cases complexity has
increased over time in the history of the universe. David Wolpert (2013)
offers a way to use the Second Law of Thermodynamics to explain
complexity by measuring it with a notion of self-similarity at different
scales.
The relationships between thermodynamic concepts, such as entropy and
free energy, and complexity are tantalising, but so far there is no agreement
about what exactly they are. The advantage of such views is that there does
indeed seem to be an important connection between free energy and
complexity in many systems and that real complex systems are always open
systems (as discussed in Chapters 2 and 3) (Smith and Morowitz 2016).
One disadvantage is that the notion of free energy is defined in terms of the
thermodynamic concepts of entropy and heat, which are well-defined in
equilibrium thermodynamics only. The analysis of the preceding chapters
shows that being open or driven is not sufficient for all the features of
complex systems.
Complexity as Self-Organised Criticality
According to Per Bak (1996) the concept of self-organised criticality could
be taken as the basis for a theory of complexity. The sandpile model shows
that complex behaviour in the form of avalanches and the formation of
structure can have universal aspects that arise spontaneously from simple
interactions of many parts. Such self-organisation occurs only in open
systems. It is a kind of emergence not exhibited by gases and condensed
matter away from critical points but which falls well short of the more
sophisticated features of complex systems involving various forms of
adaptive behaviour. Hence, while it is an important feature of complex
systems (see Chapter 4), it is not the only one.
5.3.3 Computational Conceptions of Complexity
Recall that the influential idea of ‘logical depth’ due to computer scientist
Charles Bennett (1991) is based on the mathematical theory of
computational complexity, which measures the degree to which strings of
symbols depart from randomness (as is explained in Chapter 4). Physicist
Seth Lloyd (2006) argues that complexity results from quantum
computation, assuming both the existence of quantum multiverses and that
complexity is measured by the notion of thermodynamic depth. Many other
computational and information-theoretic measures of complexity have been
proposed. Chapter 4 shows they measure not complexity but order. There is
a rivalry between physical and computational and information-theoretic
measures of complexity throughout debates about complexity science. The
next section is about conceptions of complexity that one way or another
involve the idea of function.
5.3.4 Functional Conceptions of Complexity
Chief among the examples of complex systems discussed in this book and
more generally are living organisms, their parts, and collections of them –
for example, cells, eusocial insect colonies, ecosystems, the immune system
and the human brain. In the context of biology, complexity involves not
only structure and its formation and maintenance, but also adaptive
behaviour. As argued in Chapter 3, adaptive behaviour generates new forms
of features of complex systems such as robustness and nested structure. For
this reason, as mentioned in Chapter 1 and at the end of Chapter 3, many of
those who have addressed the question of what is a complex system say that
complex systems are those that display adaptive behaviour. Prominent
examples include Mitchell (2011) and Holland (2014) (and this view is
represented in the quote from Pines in Chapter 1). Biological complexity is
central to Herbert Simon’s famous idea of the hierarchy of complexity
discussed in Chapter 3.
However, the line between the living and nonliving is not the same as the
line between things that display adaptive behaviour and those that do not.
Adaptation is always relative to some notion of function, whether this be
simply the goal of reproduction or a derivative goal like navigation. All
living systems can be understood in functional terms; however, there are
also nonliving complex systems that have inherited goals and purposes
from us. Adaptation exists in artificial systems that are not alive, such as in
software. The complex systems of human construction such as markets and
economies, IT networks, transportation networks and cities can all be
thought of in functional terms – for example, Internet routers that optimise
the flow of data. Much of complexity science is the study of adaptive
behaviour in living systems or in those we have engineered, and it could be
argued that all genuinely complex systems are either living systems or
systems created to serve the purposes of living systems. Nonetheless, the
relevant concept is not that of life but that of function (or relatedly, goal or
purpose).
It may seem obvious that, given the difference between earthworms and
human beings, biological complexity has increased more or less
monotonically during the course of the history of life on Earth; however
Stephen Jay Gould argues that this is based on a very simplistic
understanding of evolution (see, for example, Gould 2011). Although
biological complexity is much discussed by influential figures such as
Simon Conway-Morris and Stuart Kauffman, there is no general agreement
about what it is. Proposed measures of it include genome length and the
subtler notion of generative entrenchment. Like the idea of logical depth
mentioned above, biological complexity is often related to the history of a
system or its relation to its environment. Physicist Eric Smith (2013)
analyses biological complexity in terms of the notions of memory and
robustness that often figure in the science of complex systems. A flock of
birds dispersed by a predator reforms and returns to its trajectory, despite
lacking any overall controller. Smith connects this with the formalism of
order parameters in statistical physics mentioned in Chapter 2. On the other
hand, biologist David Krakauer (2013) thinks of biological complexity in
terms of the cognitive capacity of an organism to represent and predict its
environment.
None of the above accounts of complexity and complex systems is
complete. Biological conceptions do not apply to nonliving systems, and
functional conceptions more generally do not apply to systems that lack
function, goals or purposes like the BZ reaction and the solar system.
However, as shown in the previous chapters, such nonliving systems exhibit
many of the features of complex systems. None of the other conceptions of
complexity discussed above are satisfactory on their own, because they
either depend on the speculative extension of thermodynamic or
computational/information-theoretic concepts beyond their domains of
application or are based on measures of one aspect of complex systems,
such as the order that they produce or their robustness, while neglecting
others. Chapter 4 discussed in detail some of the many measures of
complexity that have been proposed. If we are right that none of them
measures complexity as such, but rather features of complex systems, then
realism about complex systems based on any of them is not viable. In the
next section we explain our answer to the question ‘What is a complex
system?’ The view advocated may be thought of as a version of realism that
takes the conceptions of complexity above to be expressing different
features of complex systems so that there are different kinds of complex
systems, both because not all have all the features and because the different
features take different forms.
5.4 The Varieties of Complex Systems
Complexity science studies how real systems behave. The models of the
traditional sciences often treat systems as closed. Real complex systems
interact with an environment and have histories. Complexity is not a single
phenomenon but the features of complex systems identified in Chapter 3 are
common to many systems. If it is right that the hallmark of complex
systems is emergence and that there are different kinds of emergent features
of complex systems, then instead of defining a complex system in terms of
one particular kind of emergence, it is possible to identify different varieties
of complex systems according to what emergent features they exemplify.
This view is explained below, and the most important lessons of this book
are revisited and brought together.
The truisms of Chapter 1 fit with our analysis. As explained at the end of
Chapter 1, complexity science began with the search for new syntheses
between models and theories from different domains. For this reason it
involves multiple disciplines. Complexity science is possible because
despite the differences between different complex systems, there are kinds
of invariance and forms of universal behaviour in complex systems (truism
5), and many of these can be described mathematically, as shown in
Chapter 4. The new kinds of universality and invariance that emerge when
complex systems are modelled as networks and information-processing
systems include scaling laws and some of the forms of robustness and
modularity explained in Chapter 4.
Simple and regular behaviour of the whole can emerge from very
complicated and messy underlying behaviour of the parts, as with the
dynamics of some physical systems like the periodicity of a chemical
oscillator. On the other hand, the adaptive behaviour of an ant colony can be
generated by relatively simple rules governing individuals. Clearly,
‘simplicity’ means many different things in the context of complex systems,
but the features of complex systems can arise from relative simplicity
because of the numerosity of interactions and external drivers.
Complexity science is probabilistic in part because it uses ideas and
techniques from statistical mechanics, and in part because it models systems
in terms of information and order which is represented with probability
theory. Complexity science is computational in two ways. First, it uses
computational models of systems such as condensed matter, the solar
system, the climate and so on. Second, it models systems as themselves
performing computations. This requires a functional understanding of the
system.
Complexity science does study something distinctive – namely the
emergent features of systems that are composed of a lot of components that
interact repeatedly in a disordered way. The reason why it has been hard to
identify what is distinctive about complex systems is that there are many
different kinds of emergent properties and products of complex systems,
and they are not all found in all complex systems. The common features of
complex systems manifest themselves differently in different kinds of
systems. For example, the robustness of an insect colony is different from
the robustness of a weather pattern. It is essential also to distinguish the
order that complex systems produce and the order of complex systems
themselves.
Emergence often takes the form of relatively sharp boundaries in time,
space or both. The examples from the physical world show that these sharp
boundaries (like thermoclines and the heliopause in the solar system) and
emergent entities and properties (like gases with their pressure) are relative
to scales of energy, space and time. Section 2.1 shows that much of physics
is about emergent phenomena and that the chemistry and physics of
nonliving systems exemplifies all of the other truisms of complexity
science. The fact that nonliving systems can generate order is exemplified
by matter, the Earth, and the solar system.
Recall that order, structure or form is inhomogeneity of some kind.
Hence, in general, order is associated with symmetry breaking of some
kind. There is a general idea of equilibrium states as states that do not
change in some relevant respect over some relevant time scales. In some
living systems there is a very narrow range of temperature that allows for
the stability of the physical processes that sustain life. Phase transitions are
emergent order in the dynamics of systems of many parts, and they may
also produce structure in the systems with which they interact. Both
symmetry breaking and phase transitions are found throughout physics from
cosmology to quantum field theory. Universal aspects to critical phenomena
are found in many different contexts and in many very different kinds of
processes.
An isolated gas at thermal equilibrium does not exhibit any feedback or
self-organisation, and it does not have any of the products of complex
systems other than the emergence of the most basic kind of order (in the
form of the ideal gas laws). All complex systems that have any kind of self-
organisation beyond this are open to the environment and driven from
outside in some way. Driven physical systems like condensed matter or
gases undergoing heating across a critical point exhibit all the rich structure
of phase transitions, and these involve highly nonlinear dependencies
between their degrees of freedom.
The related phenomenon of a simple physical system self-organising
close to but without crossing a critical point has led to the term ‘self-
organised criticality’, which was discussed above and in Chapter 4. The
canonical example of this phenomenon is the pile of sand, which originally
was only a computer algorithm. While systems at phase transitions exhibit
order and nonlinearity (in the form of power-law behaviour), self-organised
critical systems represent a higher level of complexity since they also
exhibit robustness as they stay close to a critical point under a range of
external conditions exhibiting correlations on many length scales. Another
example of a driven system is the BZ reaction, which spontaneously
produces order and obeys nonlinear emergent dynamics. These are complex
systems of the most simple kind. A much bigger and more complex purely
physical system like the solar system, on the other hand, has further
features, including history and nested structure, as well as robustness of
various kinds.
Clearly, more is different in many different ways, and the ‘more’ in
question can be parts, interactions, connections, and so on. There are many
different kinds of emergence, and not all complex systems exemplify all of
them. The full extent of the emergence of the world around us did not
happen overnight. Even the most simple nonliving things with which we
interact, such as stones and the wind, exist only because of a very long and
complex history. This is even more true for living beings, which carry a lot
of the history of life within them, as well as the history of the matter of
which they are made. In multi-cellular life in general, and especially in
humans and our brains, function is built on function so that history is
encoded in both mechanisms and structure.
The examples from the physical world are sufficient to show that there
are many forms of structure, at many different length and time scales and
that there are many examples of relatively separate slow and fast dynamics
within a single system in physics. Processes on different time scales can
become effectively decoupled. Indeed, physics is full of precise laws about
relatively isolated systems and their different kinds of properties at many
different scales because of the effective decoupling of scales. Biology is not
like this because processes on very different scales are tightly coupled in
organisms.
Decoupled fast and slow dynamics can produce and maintain the
stability of structure, as within atoms and the solar system. The climate is a
case where the structure of the system and the structure produced by the
system are intricately related. In the atmosphere there are interconnected
processes on many different length and time scales and with both positive
and negative feedback loops. The climate and the economy illustrate how
complex systems can go wrong when processes at different scales become
tightly coupled. Complex systems often have very different equilibria in
which the very same collection of parts displays very different emergent
properties. For example, a stable weather pattern in the form of seasonal
rains can disappear, and stable prices can become hyperinflation.
In general, positive feedback is often destabilising and negative feedback
stabilising. The discussion of markets and economies shows that the effects
of feedback can be highly unpredictable. In general, the role of feedback in
the interaction between people and technology is extremely important. For
example, positive feedback loops have led to the dominance of very large
service providers and to people mediating more and more of their
interactions with each other by smartphones. The next most important
feedback effect will be between human beings and intelligent software
agents and big data (Burr et al. 2018).
It is stochasticity and feedback that generate higher-level approximate
order. Only when there are many instances of probabilistic processes does
average behaviour dominate what happens. Hence, the kind of robustness to
perturbation found in complex systems requires large numbers. This is
another way ‘more is different’. Larger numbers are needed for interactions
between parts to become frequent enough for self-organisation to occur.
This is yet another variation on ‘more is different’. It is important that the
idea of interaction here is the general one of dependence of the states of the
elements on each other, which may be mediated by different kinds of
physical interactions.
As noted in Chapter 1 and above, the disagreement about how to define
complexity goes along with disagreement about which systems are
complex. Nobody denies the brain is a complex system, but the solar
system is more contentious. Certainly, the brain is the complex system that
exhibits all features of complexity to the highest degree. However, to say
that the BZ reaction is not a complex system but the brain is misses the
point that they share the basic conditions of complex systems and some of
the important products, even though they do not both have all of the latter.
Complexity is a multi-faceted phenomenon that has a variety of features
not all of which are found in all complex systems but which are related, as
when feedback produces nonlinearity. The ‘complexity’ measures
applicable to real-world complex systems do not measure complexity as
such but rather a variety of features, each of which can take different forms.
Hence, rather than restricting the term ‘complex system’ to systems
displaying adaptive behaviour or using it completely generally as if all
complex systems were simply those displaying some kind of self-
organisation, our view is that there are different kinds of complex systems
and that different features of complexity are displayed by them. The
‘conditions’ for complexity – numerosity, disorder and diversity, feedback,
and non-equilibrium – give rise to the ‘products’ of order and organisation,
robustness, and nonlinearity. Figure 5.1 shows a sketch of these features of
complexity. Systems with all the features of order, robustness, nonlinearity,
nested structure and history are at the highest level of complexity found in
systems that are not ascribed functions. Once function is involved then
adaptive behaviour leads to the functional features of forms of self-
organisation and order, modularity, and memory – which are interlinked and
which require all of the other conditions and products.
However, there are also more or less sophisticated forms of adaptive
behaviour, and one might assign a degree of adaptive behaviour to the
Earth’s climate. This explains why what people mean by ‘complexity’ and
‘complex system’ is so fluid. They are sometimes talking about some or all
of the products, sometimes about the conditions, and sometimes about both.
The climate is the biggest system we know of at that level of complexity.
But the climate also illustrates how permeable the boundaries are between
the different levels of complexity. The Earth’s atmosphere developed
because of the simultaneous evolution of microbes, exemplifying the
intricate link between living and nonliving matter. One might even argue
that the climate is performing a function, that of making life on Earth
possible in its synergy between living organisms and lifeless matter. (The
Gaia hypothesis makes this view explicit; Lovelock and Margulis 1974.)
It is fundamental to all complex systems that have all the features above
that they have a very long history and that their interesting features are
apparent only at the right spatial and temporal scales. The stability of
complex systems arises spontaneously only under the right conditions. If
they are perturbed too much, the products of them that we value can
dissolve into the disorder that underlies them. Interference with them can
shift them into new states or into states in which complexity no longer
emerges.
In answer to the questions at the beginning of the preface, complexity is
not just a label, but it is not a single phenomenon. The different conceptions
of complexity of physicists, biologists, social scientists and others cannot be
brought into a single framework, because the features of complex systems
are many and take different forms. However, there are multiple frameworks,
such as information theory, network theory, and the theory of critical
phenomena, that can be applied across the sciences. Measures of
complexity are meaningful but measure different features of complex
systems that manifest themselves in different ways.
Figure 5.1: The products of complexity, separated into those that appear in matter, functional systems
and living systems and those that appear in functional and living systems only.
5.5 Implications
There are two broad kinds of implications of this book – namely the
philosophical and the practical – and they shade into one another when the
ethics and politics of complexity science are considered. There are many
important theoretical questions on which complexity science bears, the
most obvious ones concerned with the relationships between life and
nonliving matter, and between conscious and non-conscious matter. The
general implication of our analysis for these matters is that the dichotomy
between atoms and molecules and advanced life forms is a very crude way
of seeing the many layers of structure that are found at different scales. The
only way to understand the emergence of life is by studying the processes
that occur in self-organising physical systems not just physical structures.
Once the complexity of nonliving systems, such as the solar system and the
Earth and its climate, is grasped in detail, the difference between life and
non-life seems to be less of a mysterious leap and more of a continuum.
Similarly, the intelligent collective behaviour of eusocial insects and the
many layers of complexity between the simplest nervous systems and the
human brain make the consciousness of people (who of course have been
subject to the very many interactions of development) seem like a very
complicated bundle of capacities, not a single property. More generally, one
of the main metaphysical lessons of this book is that emergence comes in
many different forms at many different scales and should be understood in
terms of interactions and processes.
One of the most important practical goals of complexity science is to
improve the functioning of societal processes that make important
decisions. An important lesson of complexity science is that complex
systems can spontaneously produce emergent phenomena in some regimes
but not in others. There are many social processes at work in the scientific
community, which itself may be thought of as a complex system. As such,
at its best, it has the emergent properties of error correction and truth
tracking and produces incredibly accurate and precise information about the
world. However, complex systems can be pushed into very different
equilibria in which they completely lose some or all of their emergent
properties. Care must be taken to ensure that the incentives and forms of
interactions that scientists are subject to continue to produce science as a
collective result. Clearly, the complex systems of the climate and
ecosystems may be destabilised to the point of radical change due to the
coupling of processes at different time scales and to various kinds of
perturbations that may not be obviously significant in advance. The
dramatic changes currently being witnessed may be only the tip of the
iceberg.
Recall the example of the army ants and the need for sufficient numbers
for their interactions to produce collective behaviour. Hence, collective
human behaviour may be similarly sensitive to the nature and number of
interactions between individuals. As technology is increasingly eliminating
or mediating these interactions, it is possible that human beings and human
society will change radically in ways that cannot be anticipated. Positive
feedback loops can act very forcefully and fast, as anyone who has heard
the sound from a microphone fed back into it from loudspeakers knows.
The sound volume goes up increasingly fast until someone or something
cuts the connection. There are many positive feedback loops operative at
the moment as the use of different technologies goes from being exotic to
optional to effectively compulsory for individuals.
The fields of AI and Big Data are about complex systems. The old way
to think about artificial intelligence involved a computer program that was
designed in advance and didn’t change when in use. The systems that are
now used to solve the classic problems such as face recognition and
translation are open systems that have interacted extensively before they are
used and continue to do so once they are in use. It is arguable that new
complex systems are being created whose emergent behaviour is
completely unpredictable and some of which may not be desirable. For
example, the complex system that consists of social media users and the
algorithms that select the content that is most circulated have created
emergent phenomena such as ‘trending’ and can allow antisocial attitudes
and misinformation to propagate more effectively that prosocial attitudes or
facts respectively.
In general, since the equilibria that dominate the natural and social
worlds have evolved over millennia to be compatible with human
flourishing, it is much more likely that any new equilibria into which they
shift will be to our detriment, and it is possible that they will lead to the
destruction of much or all of what we value. The universe seems to have
become more and more complicated since its infancy, and the most intricate
structures in the universe of which we know (which are living systems)
seem to be more complex than they have ever been. However, there is no
guarantee that this will continue, and the fate of complex systems such as
the climate, the economy and life on Earth depends on us.
1Similarly, computer scientists do not need to worry about what a computer is. However, Alan
Turing’s investigations into the latter question led to the theory of Turing machines, which is a
valuable part of computer science.
2Some philosophers deny that there is ontological emergence and claim that anything other than
fundamental physical stuff is at best a kind of second-class thing (see Ney 2014, pp. 111–112). We
reject this view.
Appendix – Some Mathematical
Background
In this Appendix, all mathematical terminology used in the main text is
defined, and some more background is given to the mathematical
formalisms. The areas which are included are probability theory,
information theory, algorithmic complexity, and network theory. All
terminology which is defined is typeset in bold.
A Probability Theory
An alphabet is a set of symbols, numeric or symbolic, continuous or
discrete, finite or infinite. Symbolic alphabets are discrete; continuous
alphabets are numeric and infinite. | | denotes the size of set . An
example of a numeric finite, discrete alphabet is the binary set {0,1}; an
example of a symbolic, finite alphabet is a set of Roman letters {a,b,c,...,z}.
An example of a numeric infinite, discrete alphabet is that of the natural
numbers N, an example of a continuous alphabet is the set of real numbers
R. This book discusses only discrete alphabets.
A discrete random variable X is a discrete alphabet equipped with
a probability distribution P(X) ≡ {Pr(X = x), x ∈ }. We denote the
probabilities Pr(X = x) by P(x) or sometimes, to avoid confusion, by PX (x).
The uniform distribution of a set is the distribution P(x) = 1/| | for
all x ∈ . For two discrete random variables, X and Y, the joint
probabilities Pr(X = x,Y = y) on alphabet × are denoted by P(xy) or
sometimes, to avoid confusion, by PXY (xy). The joint probability
distribution induces a conditional probability distribution P(x|y) ≡ Pr(X =
x|Y = y), which is a probability distribution on conditioned on Y taking
particular value Y = y. Any joint probability P(xy) can be written as
The expectation value of a discrete numeric random variable X, denoted
〉
by X , is defined as
Another common notation for the expectation value of X is .
The variance of a numeric random variable X is the average deviation of
X from its expectation value. Denoted by VarX, it is defined as
The square root of the variance of a random variable X is called the
standard deviation σ:
The ratio of variance to expectation value is called the coefficient of
variation,
The covariance of two numeric random variables X and Y is defined as
A stochastic process {Xt}t∈T is a sequence of random variables Xt,
defined on a joint probability space, taking values in a common set ,
indexed by a set T which is often or and thought of as time. This book
only discusses discrete time processes. A stochastic process is called a
Markov chain if Xt (sometimes called ‘the future’) is probabilistically
independent of X0...Xt−2 (‘the past’), given Xn−1 (‘the present’); in other
words,
A stochastic process is stationary if
A hidden Markov model {Xt,Yt}t∈T is a stationary stochastic process of
two random variables Xt and Yt which forms a Markov chain in the sense
that Yt depends only on Xt, and Xt depends only on Yt−1 and Xt−1:
The graphical representation of a hidden Markov model is a directed
graph where the states are the outcomes x ∈ of the random variable Xt
and the state transitions are labelled by the outcomes y ∈ of the random
variable Yt and the corresponding conditional probability P(Yt+1 = y,Xt+1 =
x|Xt = x).
B Shannon Information Theory
In the 1940s, the American engineer Claude Shannon, working for Bell
Labs, introduced a mathematical theory of communication that is now at the
heart of every digital communication protocol and technology, from mobile
phones to e-mail encryption services and wireless networks (Shannon
1948). Shannon was concerned with defining and measuring the amount of
information communicated by a message transmitted over a noisy telegraph
line. He saw a message as communicating information if the receiver of the
message could not predict with certainty which message out of a set of
possible ones she would receive. By setting the amount of information
communicated by a message x as proportional to its inverse log probability
1/log P(x), Shannon axiomatically derived a measure of information, now
called Shannon entropy. The Shannon entropy, a function of a probability
distribution P but often written as a function of a random variable X, is
defined as follows (Cover and Thomas 2006):
where the log is usually base 2 and 0log0 := 0. The equivalent definition
where P = {p1, p2,..., pn}, makes it explicit that H is a function of the
probabilities alone, independent of the alphabet . This book discusses
only the entropy of finite probability distributions, but the definition of the
Shannon entropy extends to infinite but discrete, as well as to continuous
probability distributions. Taking the logarithm to base 2 is a convention
dating back to Shannon, due to a bit being the essential unit of computation.
For a given set of messages , the Shannon entropy is maximum for the
uniform distribution and proportional to the logarithm of the total number
of messages. This illustrates that the Shannon entropy is a measure of
randomness. If one of the messages has probability 1 and the others have
probability 0, then the message is perfectly predictable, and the Shannon
entropy is zero. The Shannon entropy is also precisely the expectation value
of the function 1/log P(x).
The joint entropy of n random variables X1,...,Xn with joint probability
distribution P(X1X2 ...Xn) is defined as
Consider two random variables X and Y and joint probability distribution
PXY. The conditional entropy of X given Y, H(X|Y), is defined as
The entropy rate of a stochastic process {Xt}t∈T is defined as
A different definition of entropy rate is as follows:
For stationary stochastic processes, h = h′. The entropy rate H(Xn|X1
...Xn−1), for finite n, is denoted by hn.
Shannon introduced the mutual information as a measure of correlation
between two random variables X and Y, defined as follows:
The mutual information is a measure of the predictability of one random
variable when the outcome of the other is known. Note that the mutual
information is symmetric in its arguments and hence measures the amount
of information ‘shared’ by the two variables. The mutual information is a
general correlation function for two random variables, measuring both
linear and non-linear correlations. In contrast to the covariance and many
other correlation measures, it is also applicable to non-numeric random
variables such as the distribution of words in an English text or the
distribution of amino acids in a DNA sequence. This is one reason why it is
widely used in complex systems research.
C Algorithmic Information Theory
A mathematical formalisation of randomness and information without
reference to probabilities was developed independently by the Soviet
mathematician Andrey Kolmogorov and the American mathematicians Ray
J. Solomonoff and Gregory Chaitin in the 1960s. They considered
information as a property of a single message, rather than of a set of
messages and their probabilities. A message is a string of letters from an
alphabet, such as the Roman alphabet or the binary characters 0 and 1. An
example of a string is ‘Hello, World!’. The string is composed of letters
from the Roman alphabet and from a set containing the comma and space
characters and the exclamation mark.
The algorithmic information content of a string is, roughly speaking,
the length of the shortest computer program which outputs the string. For
the string ‘Hello, World!’, this is probably a program of the form
‘print(“Hello, World!”)’ which has roughly the same length as the string
itself. However, for a string of 10,000 zeros and ones alternating, the
shortest program is shorter than the string itself, and the string is called
‘compressible’. The notion of compressibility is meaningful only with
longer strings. Only perfectly random strings are completely
incompressible, therefore algorithmic information is a measure of
randomness. It can be confusing that the term ‘information’ is used for
randomness, but one may think of randomness as the amount of information
which has to be communicated to reproduce the string ‘exactly’,
irrespective of how interesting the string is in other respects.
The precise definition of algorithmic information is as follows (Li and
Vitányi 2009). Consider a string x, a computing device and programs p
of length l(p). The algorithmic information of the string, K(x), is the length
of the shortest program p, which, when fed into a machine , produces
output x, (p) = x,
The minimisation is done over all possible programs. There is a
fundamental problem with carrying out the minimisation procedure:
whether an arbitrary program will finish or run forever cannot be known in
general. This is called the halting problem. It is one of the deepest results in
computer science, and due to the British mathematician Alan Turing. As a
consequence, the algorithmic information is not computable in principle,
though it can often be approximated in practice. Other names for
algorithmic information are ‘algorithmic complexity’ or ‘Kolmogorov
complexity’.
The fundamental insight of Kolmogorov, Solomonoff and Chaitin is that
the minimum length of a program is independent of the computing device
on which it is run (up to some constant which is independent of the string).
Hence, the definition of algorithmic information refers to a universal
computer, which is a fundamental notion introduced by Alan Turing in the
1940s. Algorithmic information is therefore a ‘universal’ notion of
randomness for strings because it is context- (machine-) independent. On
the other hand, the Shannon entropy is context-dependent, since it may
assign different amounts of information to the same string when it is
embedded in different sets with different probabilities.
The length is not the only important parameter of a program; its running
time is of equal importance. There are very short programs that take a long
time to run, while the print program might be long but finished very
quickly. This trade-off is relevant to the measures of complexity (discussed
at the end of Chapter 4), which include the well-known logical depth.
D Complex Networks
A network, or a graph, is a set of nodes and, for simplicity here, there is at
most one edge between any ordered pair of nodes. Nodes and edges are also
called vertices and links, respectively. In a directed network each edge has
a directionality, beginning at one node and ending at another. In an
undirected network there is no such distinction between the start and end
node of an edge. An example of an undirected network is the Internet. The
servers are nodes, and edges between them are the physical wirings. An
example of a directed network is an ecological food web. Two animals are
linked if one of them feeds on the other so that a predator has a directed
edge to its prey.
The degree of a node in a network is the number of edges attached to it.
In a directed network, one distinguishes between in-degree and out-degree.
The in-degree of a node is the number of edges directed to the node, and the
out-degree is the number of nodes directed away from it.
Mathematically, a network of n nodes is represented by an adjacency
matrix, A, which is an n × n matrix where each non-zero entry Aij
represents an edge from node i to node j (Newman 2010). In an unweighted
network the Aij are 1 if an edge exists from node i to node j and 0 otherwise.
A weighted network assigns a real number to each edge, Aij ∈ . Such
weights could, for example, represent the volume of data traffic between
two servers. In an undirected network, Aij = Aji since Aij and Aji refer to the
same object. The in-degree of a node i is the number of non-zero entries in
the jth column of A. The out-degree of a node i is the number of non-zero
entries in the ith row of A. In an undirected network, these numbers are
equal.
The degree distribution of a network is the frequency distribution over
node degrees. A uniform degree distribution, for example, means that nodes
of degree 1 are equally likely as nodes of degree n. A path is a sequence of
nodes such that every two consecutive nodes in the sequence are connected
by an edge. In a directed network the nodes have to be connected by edges
that all point in the forward direction. The path length is the number of
edges traversed along the sequence of a path. The shortest path between
two nodes is the sequence with the minimum number of traversed edges to
get from one node to the other. The average shortest path is the sum of all
shortest path lengths divided by their number. The diameter of a network is
the longest of all shortest paths.
Bibliography
L. A. Adamic and B. A. Huberman. Power-law distribution of the world
wide web. Science, 287(5461):2115, 2000.
R. Albert and A.-L. Barabási. Statistical mechanics of complex networks.
Reviews of Modern Physics, 74(1):47–97, 2002.
R. Albert, H. Jeong, and A. L. Barabási. Diameter of the world-wide web.
Nature, 401(6749):130, 1999.
R. Albert, H. Jeong, and A.-L. Barabási. Error and attack tolerance of
complex networks. Nature, 406(6794):378–382, 2000.
Michel Anctil. Dawn of the neuron: The early struggles to trace the origin
of nervous systems. McGill-Queen’s Press, 2015.
P. W. Anderson. More is different. Science, 177(4047):393–396, 1972.
W. B. Arthur. Complexity and the economy. Science, 284(5411):107–109,
1999.
J.-P. Aubin. A survey of viability theory. SIAM Journal on Control and
Optimization, 28(4):749–788, 1990.
Jean-Pierre Aubin. Viability Theory. Springer Science & Business Media,
2009.
C. Aymanns, J. D. Farmer, A. M. Kleinnijenhuis, and T. Wetzer. Models of
financial stability and their application in stress tests. Handbook of
Computational Economics, 4:329–391, 2018.
F. A. C. Azevedo et al. Equal numbers of neuronal and nonneuronal cells
make the human brain an isometrically scaled-up primate brain. Journal
of Comparative Neurology, 513(5):532–541, 2009.
P. Bak, C. Tang, and K. Wiesenfeld. Self-organized criticality. Physical
Review A, 38(1):364, 1988.
Per Bak. How Nature Works: The Science of Self-Organised Criticality.
Copernicus Press, 1996.
A.-L. Barabási and R. Albert. Emergence of scaling in random networks.
Science, 286(5439):509–512, 1999.
M. Baringer and J. C. Larsen. Sixteen years of Florida current transport at
27°N. Geophysical Research Letters, 28(16):3179–3182, 2001.
S. Battiston et al. Complexity theory and financial regulation. Science, 351
(6275):818–819, 2016.
Eric D. Beinhocker. The Origin of Wealth: Evolution, Complexity, and the
Radical Remaking of economics. Harvard Business Press, 2006.
A. Bekker et al. Dating the rise of atmospheric oxygen. Nature, 427(6970):
117–120, 2004.
C. Béné and L. Doyen. From resistance to transformation: A generic metric
of resilience through viability. Earth’s Future, 6(7):979–996, 2018.
L. Benedetti-Cecchi, L. Tamburello, E. Maggi, and F. Bulleri. Experimental
perturbations modify the performance of early warning indicators of
regime shift. Current Biology, 25(14):1867–1872, 2015.
C. H. Bennett. Logical depth and physical complexity. In Rolf Herken,
editor, The Universal Turing Machine – a Half-Century Survey, pages
227–257. Oxford University Press, 1991.
T. Berners-Lee, R. Cailliau, J.-F. Groff, and B. Pollermann. World-wide
web: The information universe. Internet Research, 20(4):461–471, 2010.
Ludwig von Bertalanffy. General System Theory: Foundations,
Development, Applications. George Braziller, 1969.
L. M. A. Bettencourt. The origins of scaling in cities. Science, 340(6139):
1438–1441, 2013.
L. M. A. Bettencourt, J. Lobo, D. Helbing, C. Kühnert, and G. B. West.
Growth, innovation, scaling, and the pace of life in cities. Proceedings of
the National Academy of Sciences, 104(17):7301–7306, 2007.
W. Bialek, I. Nemenman, and N. Tishby. Predictability, complexity, and
learning. Neural Computation, 13(11):2409–2463, 2001.
W. Bialek et al. Statistical mechanics for natural flocks of birds.
Proceedings of the National Academy of Sciences, 109(13):4786–4791,
2012.
James Binney. The Theory of critical phenomena : an introduction to the
renormalization group. Clarendon Press, 1992.
H. M. Blalock, editor. Causal Models in the Social Sciences. Routledge,
1985.
U. Brandes et al. On modularity clustering. IEEE Transactions on
Knowledge and Data Engineering, 20(2):172–188, 2007.
K. H. Britten, M. N. Shadlen, W. T. Newsome, and J. A. Movshon.
Responses of neurons in macaque MT to stochastic motion signals.
Visual Neuroscience, 10(6):1157–1169, 1993.
W. Buffett. Annual report, 2002.
http://www.berkshirehathaway.com/2002ar/2002ar.pdf.
E. Bullmore and O. Sporns. Complex brain networks: Graph theoretical
analysis of structural and functional systems. Nature Reviews
Neuroscience, 10(3):186–198, 2009.
E. Bullmore and O. Sporns. The economy of brain network organization.
Nature Reviews Neuroscience, 13(5):336, 2012.
C. Burr, N. Cristianini, and J. Ladyman. An analysis of the interaction
between intelligent software agents and human users. Minds and
Machines, 28(4):735–774, 2018.
J. Butterfield. Emergence, reduction and supervenience: A varied
landscape. Foundations of Physics, 41(6):920–959, 2011a.
J. Butterfield. Less is different: Emergence and reduction reconciled.
Foundations of Physics, 41(6):1065–1135, 2011b.
S. R. Carpenter et al. Early warnings of regime shifts: A whole-ecosystem
experiment. Science, 332(6033):1079–1082, 2011.
E. Chaisson. Using complexity science to search for unity in the natural
sciences. In C. H. Lineweaver, editor, Complexity and the Arrow of Time.
Cambridge University Press, 2013.
Austin Chambers. Modern Vacuum Physics. CRC Press, 2004.
A. Clark. Whatever next? predictive brains, situated agents, and the future
of cognitive science. Behavioral and Brain Sciences, 36(3):181–204,
2013.
P. U. Clark, N. G. Pisias, T. F. Stocker, and A. J. Weaver. The role of the
thermohaline circulation in abrupt climate change. Nature, 415(6874):
863–869, 2002.
A. Clauset, C. Moore, and M. Newman. Hierarchical structure and the
prediction of missing links in networks. Nature, 453(7191):98–101,
2008.
A. Clauset, C. R. Shalizi, and M. Newman. Power-law distributions in
empirical data. SIAM Review, 51(4):661–703, 2009.
Computational Mechanics Group, 2015. http://cmpy.csc.ucdavis.edu/.
G. Coricelli et al. Regret and its avoidance: A neuroimaging study of choice
behavior. Nature Neuroscience, 8(9), 2005.
I. D. Couzin and N. R. Franks. Self-organized lane formation and optimized
traffic flow in army ants. Proceedings of the Royal Society of London B:
Biological Sciences, 270(1511):139–146, 2003.
I. D. Couzin and J. Krause. Collective memory and spatial sorting in animal
groups. Theoretical Biology, 218, 2002.
Thomas M. Cover and Joy A. Thomas. Elements of Information Theory.
Wiley-Blackwell, 2nd edition, 2006.
Credit Suisse Research Institute. Global wealth databook. Technical report,
2016.
J. P. Crutchfield. The calculi of emergence: Computation, dynamics and
induction. Physica D: Nonlinear Phenomena, 75(1-3):11–54, 1994.
J. P. Crutchfield. Between order and chaos. Nature Physics, 8(1):17, 2012.
J. P. Crutchfield and D. P. Feldman. Regularities unseen, randomness
observed: Levels of entropy convergence. Chaos: An Interdisciplinary
Journal of Nonlinear Science, 13(1):25–54, 2003.
J. P. Crutchfield and C. R. Shalizi. Thermodynamic depth of causal states:
Objective complexity via minimal representations. Physical Review E,
59 (1):275, 1999.
J. P. Crutchfield and K. Young. Inferring statistical complexity. Physical
Review Letters, 63(2):105, 1989.
J. P Crutchfield, C. J. Ellison, and J. R. Mahoney. Time’s barbed arrow:
Irreversibility, crypticity, and stored information. Physical Review
Letters, 103(9):094101, 2009.
V. Dakos et al. Slowing down as an early warning signal for abrupt climate
change. Proceedings of the National Academy of Sciences, 105(38):
14308–14312, 2008.
A. Davidson. How AIG fell apart. Technical report, Reuter, 2008.
http://www.reuters.com/article/us-how-aig-fell-apart-
idUSMAR85972720080918.
P. Davies and N. H. Gregersen, editors. Information and the Nature of
Reality: From Physics to Metaphysics. Cambridge University Press,
2014.
Sybren Ruurds De Groot and Peter Mazur. Non-Equilibrium
Thermodynamics. Courier Corporation, 2013.
H. de Jong. Modeling and simulation of genetic regulatory systems: A
literature review. Journal of Computational Biology, 9(1):67–103, 2002.
J. DeFelipe, L. Alonso-Nanclares, and J. I. Arellano. Microstructure of the
neocortex: Comparative aspects. Journal of neurocytology, 31(3-5):299–
316, 2002.
Guillaume Deffuant and Nigel Gilbert. Viability and Resilience of Complex
Systems: Concepts, Methods and Case Studies from Ecology and Society.
Springer Science & Business Media, 2011.
A. J. Dunleavy, K. Wiesner, R. Yamamoto, and C. P. Royall. Mutual
information reveals multiple structural relaxation mechanisms in a model
glass former. Nature communications, 6:6089, 2015.
J. A. Dunne, R. J. Williams, and N. D. Martinez. Food-web structure and
network theory: The role of connectance and size. Proceedings of the
National Academy of Sciences, 99(20):12917–12922, 2002.
David Easley and Jon Kleinberg. Networks, Crowds, and Markets:
Reasoning About a Highly Connected World. Cambridge University
Press, 2010.
Joshua M. Epstein and Robert Axtell. Growing Artificial Societies: Social
Science from the Bottom Up. Brookings Institution Press and MIT Press,
1996.
P. Erdös and A. Rényi. On the evolution of random graphs. Publication of
the Mathematical Institute of the Hungarian Academy of Sciences, 5(1):
17–60, 1960.
A. Fabiani, F. Galimberti, S. Sanvito, and A. R. Hoelzel. Extreme polygyny
among southern elephant seals on sea lion island, falkland islands.
Behavioral Ecology, 15(6):961–969, 2004.
Kenneth Falconer. Fractal Geometry: Mathematical Foundations and
Applications. John Wiley & Sons, 2004.
P. G. Falkowski, T. Fenchel, and E. F. Delong. The microbial engines that
drive earth’s biogeochemical cycles. Science, 320(5879):1034–1039,
2008.
M. Faloutsos, P. Faloutsos, and C. Faloutsos. On power-law relationships of
the internet topology. ACM SIGCOMM Computer Communication
Review, 29(4):251–262, 1999.
E. F. Fama. Efficient capital markets: II. The Journal of Finance, 46(5):
1575–1617, 1991.
J. D. Farmer. Physicists attempt to scale the ivory towers of finance.
Computing in Science & Engineering, 1(6):26–39, 1999.
J. D. Farmer and J. Geanakoplos. The virtues and vices of equilibrium and
the future of financial economics. Complexity, 14(3):11–38, 2009.
J. D. Farmer and N. Packard, editors. Evolution, Games, and Learning:
Models for Adaptation in Machines and Nature – Proceedings of the
Fifth Annual International Conference, May 20-24, 1985, volume 22, 1-
3 1986. Physica D: Nonlinear Phenomena.
Richard P. Feynman. Feynman Lectures on Computation. Addison-Wesley,
1998.
Stanley Finger. Minds Behind the Brain: A History of the Pioneers and
Their Discoveries. Oxford University Press, 2005.
R. A. Fisher. The wave of advance of advantageous genes. Annals of
Eugenics, 7(4):355–369, 1937.
R. Foote. Mathematics and complex systems. Science, 318(5849):410–412,
2007.
S. Fortunato and M. Barthelemy. Resolution limit in community detection.
Proceedings of the National Academy of Sciences, 104(1):36–41, 2007.
N. R Franks. Army ants: a collective intelligence. American Scientist, 77:
138–145, 1989.
N. R. Franks, N. Gomez, S. Goss, and J. L. Deneubourg. The blind leading
the blind in army ant raid patterns: Testing a model of self-organization
(Hymenoptera: Formicidae). Journal of Insect Behavior, 4(5):583–607,
1991.
K. Friston. Causal modelling and brain connectivity in functional magnetic
resonance imaging. PLoS Biology, 7(2):e1000033, 2009.
S. Garnier et al. Stability and responsiveness in a self-organized living
architecture. PLoS Computational Biology, 9(3):e1002984, 2013.
M. Gell-Mann. What is complexity? Complexity, 1(1):16–19, 1995.
M. Gell-Mann and S. Lloyd. Information measures, effective complexity,
and total information. Complexity, 2(1):44–52, 1996.
Murray Gell-Mann. The Quark and the Jaguar. W. H. Freeman, 1994.
N. Goldenfeld and L. P. Kadanoff. Simple lessons from complexity.
Science, 284(5411):87–89, 1999.
Herbert Goldstein. Classical Mechanics. Addison-Wesley series in
advanced physics. Addison-Wesley, 1950.
Deborah M. Gordon. Ant Encounters: Interaction Networks and Colony
Behavior. Princeton University Press, 2010.
Stephen Jay Gould. Full House. Harvard University Press, 2011.
P. Grassberger. Toward a quantitative theory of self-generated complexity.
International Journal of Theoretical Physics, 25(9):907–938, 1986.
Richard K. Guy and John H. Conway. Winning Ways for Your Mathematical
Plays. Academic Press, 1982.
A. G. Haldane. Rethinking the financial network, 2009.
https://www.bankofengland.co.uk/speech/2009/rethinking-the-financial-
network.
A. G. Haldane and R. M. May. Systemic risk in banking ecosystems.
Nature, 469(7330):351–355, 2011.
Heiko Hamann. Swarm Robotics: A Formal Approach. Springer, 2018.
D. Haymann. How SARS was contained, 2013.
http://www.nytimes.com/2013/03/15/opinion/global/how-sars-was-
contained.html.
R. M. Hazen, D. Papineau, and W. Bleeker. Mineral evolution. American
Mineralogist, 93:1693–1720, 2008.
A. D. Henry, P. Prałat, and C.-Q. Zhang. Emergence of segregation in
evolving social networks. Proceedings of the National Academy of
Sciences, 108(21):8605–8610, 2011.
J. H. Holland. Complex adaptive systems. Daedalus, 121(1):17–30, 1992.
John H. Holland. Complexity: A Very Short Introduction. Oxford University
Press, 2014.
Bert Hölldobler and Edward O. Wilson. The Superorganism : the Beauty,
Elegance, and Strangeness of Insect Societies. W.W. Norton, 2008.
C. S. Holling. Resilience and stability of ecological systems. Annual Review
of Ecology and Systematics, 4(1):1–23, 1973.
B. Holmstrom and J. Tirole. Financial intermediation, loanable funds, and
the real sector. The Quarterly Journal of Economics, 112(3):663–691,
1997.
John E. Hopcroft, Rajeev Motwani, and Jeffrey D. Ullman. Introduction to
Automata Theory, Languages, and Computation. Addison-Wesley, 2nd
edition, 2001.
S. Houman. Credit default swaps and regulatory reform. Technical report,
Mercatus Center, George Mason University, 2009.
https://www.mercatus.org/publication/credit-default-swaps-and-
regulatory-reform.
B. A. Huberman and L. A. Adamic. Evolutionary dynamics of the world
wide web. arXiv preprint cond-mat/9901071, 1999.
Paul Humphreys. Emergence: A Philosophical Account. Oxford University
Press, 2016.
IEEE Communications Society. Infographic: The internet of things, 2015.
https://www.comsoc.org/blog/infographic-internet-things-iot.
Nobuyuki Ikeda and Shinzo Watanabe. Stochastic Differential Equations
and Diffusion Processes. North-Holland, 2014.
International Telecommunication Union. Overview of the internet of things.
Technical report, 2012. https://www.itu.int/rec/T-REC-Y.2060-201206-I.
Internet Systems Consortium. Internet host count history. Technical report,
2012. https://www.isc.org/network/survey/, Accessed August 2018.
IPCC 2013. Technical summary. In T. F. Stocker et al., editors, Climate
Change 2013: The Physical Science Basis. Contribution of Working
Group I to the Fifth Assessment Report. Cambridge University Press,
2013a.
IPCC 2013. Carbon and other biogeochemical cycles. In T. F. Stocker et al.,
editors, Climate Change 2013: The Physical Science Basis. Contribution
of Working Group I to the Fifth Assessment Report of the
Intergovernmental Panel on Climate Change. Cambridge University
Press, 2013b.
IPCC 2013. Introduction. In T.F. Stocker et al., editors, Climate Change
2013: The Physical Science Basis. Contribution of Working Group I to
the Fifth Assessment Report of the Intergovernmental Panel on Climate
Change. Cambridge University Press, 2013c.
D. Jardim-Messeder et al. Dogs have the most neurons, though not the
largest brain: Trade-off between body mass and number of neurons in
the cerebral cortex of large carnivoran species. Frontiers in
Neuroanatomy, 11:118, 2017.
E. T. Jaynes. Information theory and statistical mechanics. Physical Review,
106(4):620, 1957.
H. Jeong, S. P. Mason, A.-L. Barabási, and Z. N. Oltvai. Lethality and
centrality in protein networks. Nature, 411(6833):41, 2001.
H. Jeong et al. The large-scale organization of metabolic networks. Nature,
407(6804):651, 2000.
L. Jost. Entropy and diversity. Oikos, 113(2):363–375, 2006.
D. Kahneman. A psychological perspective on economics. American
economic review, 93(2):162–168, 2003.
J. D. Keeler and J. D. Farmer. Robust space-time intermittency and 1f
noise. Physica D: Nonlinear Phenomena, 23(1–3):413–435, 1986.
D. Kelly, M. Dillingham, A. Hudson, and K. Wiesner. A new method for
inferring hidden markov models from noisy time sequences. PloS One, 7
(1):e29703, 2012.
http://www.mathworks.com/matlabcentral/fileexchange/33217.
S.-H. Kim. Fractal structure of a white cauliflower. arXiv preprint cond-
mat/0409763, 2004.
M. Kleiber. Body size and metabolism. Hilgardia, 6(11):315–353, 1932.
A. Klein et al. Evolution of social insect polyphenism facilitated by the sex
differentiation cascade. PLoS Genetics, 12(3):e1005952, 2016.
D. Krakauer. The inferential evolution of biological complexity: Forgetting
nature by learning to nurture. In C. H. Lineweaver, editor, Complexity
and the Arrow of Time. Cambridge University Press, 2013.
J. Ladyman, J. Lambert, and K. Wiesner. What is a complex system?
European Journal for Philosophy of Science, 3(1):33–67, 2013.
S. Lawrence and C. L. Giles. Accessibility of information on the web.
Nature, 400(6740):107–109, 1999.
C.-B. Li, H. Yang, and T. Komatsuzaki. Multiscale complex network of
protein conformational fluctuations in single-molecule time series.
Proceedings of the National Academy of Sciences, 105(2):536–541,
2008.
Ming Li and Paul Vitányi. An Introduction to Kolmogorov Complexity and
Its Applications. Springer, 3rd edition, 2009.
C. Lineweaver. A simple treatment of complexity: Cosmological entropic
boundary conditions on increasing complexity. In C. H. Lineweaver,
editor, Complexity and the Arrow of Time. Cambridge University Press,
2013.
S. Lloyd. Measures of complexity: A nonexhaustive list. Control Systems
Magazine, IEEE, 21(4):7–8, 2001.
S. Lloyd and H. Pagels. Complexity as thermodynamic depth. Annals of
Physics, 188(1):186–213, 1988.
Seth Lloyd. Programming the Universe. Knopf, 2006.
E. Lorenz. Predictability: Does the flap of a butterfly’s wings in brazil set
off a tornado in texas? Presentation at American Association for the
Advancement of Science, 139th meeting, Washington, DC, 1972.
http://eaps4.mit.edu/research/Lorenz/Butterfly_1972.pdf.
J. E. Lovelock and L. Margulis. Atmospheric homeostasis by and for the
biosphere: The gaia hypothesis. Tellus, 26(1-2):2–10, 1974.
R. S. MacKay. Nonlinearity in complexity science. Nonlinearity, 21:T273,
2008.
B. F. Madore and W. L. Freedman. Computer simulations of the Belousov-
Zhabotinsky reaction. Science, 222(4624):615–616, 1983.
Klaus Mainzer. Thinking in Complexity – The Computational Dynamics of
Matter. Springer, 1st edition, 1994.
E. B. Mallon and N. R. Franks. Ants estimate area using Buffon’s needle.
Proceedings of the Royal Society of London B: Biological Sciences, 267
(1445):765–770, 2000.
Benoît B. Mandelbrot. The Fractal Geometry of Nature. W. H. Freeman,
1983.
Benoît B. Mandelbrot. Fractals and Scaling in Finance: Discontinuity,
Concentration, Risk. Springer Science & Business Media, 2013.
Benoît B. Mandelbrot and Richard Hudson. The Mis Behaviour of Markets:
A Fractal View of Risk, Ruin and Reward. Profile Books, 2010.
M. A. Marra et al. The genome sequence of the sars-associated coronavirus.
Science, 300(5624):1399–1404, 2003.
Alfred Marshall. Principles of Economics. Macmillan, 1890.
Mark Maslin. Climate: A Very Short Introduction. Oxford University Press,
2013.
H. R. Mattila and T. D. Seeley. Genetic diversity in honey bee colonies
enhances productivity and fitness. Science, 317(5836):362–364, 2007.
J. W. McAllister. Effective complexity as a measure of information content.
Philosophy of Science, 70(2):302–307, 2003.
Hans Meinhardt. Models of Biological Pattern Formation. Academic Press,
1982.
R. Menzel and M. Giurfa. Cognitive architecture of a mini-brain: The
honeybee. Trends in Cognitive Sciences, 5(2):62–71, 2001.
S. Milgram. The small world problem. Psychology today, 2(1):60–67, 1967.
Melanie Mitchell. Complexity: A Guided Tour. Oxford University Press,
2011.
M. Mitzenmacher. A brief history of generative models for power law and
lognormal distributions. Internet Mathematics, 1(2):226–251, 2004.
J. M. Montoya and R. V. Solé. Small world patterns in food webs. Journal
of Theoretical Biology, 214(3):405–412, 2002.
Moz. https://moz.com/top500/pages, 2018. accessed on 24 August 2018.
Carl D. Murray and Stanley F. Dermott. Solar System Dynamics.
Cambridge University Press, 1999.
M. Newman. The structure and function of complex networks. SIAM
Review, 45(2), 2003.
M. Newman. Power laws, Pareto distributions and Zipf’s law.
Contemporary Physics, 46(5):323–351, 2005.
M. Newman and M. Girvan. Finding and evaluating community structure in
networks. Physical review E, 69(2):026113, 2004.
M. Newman, S. Forrest, and J. Balthrop. Email networks and the spread of
computer viruses. Physical Review E, 66(3):035101, 2002.
Mark Newman. Networks: An Introduction. Oxford University Press, 1st
edition, 2010.
Alyssa Ney. Metaphysics: An Introduction. Routledge, 2014.
Grégoire Nicolis and Iiya Prigogine. Self-Organzisation in Nonequilibrium
Systems. John Wiley & Sons, 1977.
John Nolte and John W. Sundsten. The Human Brain : an Introduction to its
Functional Anatomy. Mosby, 2002.
A. Nordrum. Popular internet of things forecast of 50 billion devices by
2020 is outdated. Technical report, IEEE Spectrum, 2016.
http://spectrum.ieee.org/tech-talk/telecom/internet/popular-internet-of-
things-forecast-of-50-billion-devices-by-2020-is-outdated.
P. Nurse. Life, logic and information. Nature, 454(7203):424–426, 2008.
R. K. Pachauri, L. Meyer, and Core Writing Team. Climate change 2014:
Synthesis report. contribution of working groups I, II and III to the fifth
assessment report of the Intergovernmental Panel on Climate Change.
Technical report, 2014.
Scott E. Page. Diversity and Complexity. Princeton University Press, 1st
edition, 2010.
A. J. Palmer, C. W. Fairall, and W. A. Brewer. Complexity in the
atmosphere. IEEE Transactions on Geoscience and Remote Sensing,
38(4):2056–2063, 2000.
S. E. Palmer, O. Marre, M. J. Berry, and W. Bialek. Predictive information
in a sensory population. Proceedings of the National Academy of
Sciences, 112(22):6908–6913, 2015.
Vilfrido Pareto. Manual of Political Economy. Kelley, 1980. Reprint.
J. K. Parrish and L. Edelstein-Keshet. Complexity, pattern, and evolutionary
trade-offs in animal aggregation. Science, 284(5411):99–101, 1999.
D. Paul. Credit default swaps, the collapse of AIG and addressing the crisis
of confidence. Huffington Post, 2008.
http://www.huffingtonpost.com/david-paul/credit-default-swaps-
the_b_133891.html.
Azaria Paz. Introduction to Probabilistic Automata. Academic Press, 1971.
Roger Penrose. The Road to Reality. Jonathan Cape, 2004.
Thomas Piketty. Capital in the Twenty-First Century. Harvard University
Press, 2014.
Stuart L. Pimm. Food Webs. Springer, 1982.
D. Pines, editor. Emerging Syntheses in Science: Proceedings from the
Founding Workshops of the Santa Fe Institute. SFI Press, 2019.
D. Price. A general theory of bibliometric and other cumulative advantage
processes. Journal of the Association for Information Science and
Technology, 27(5):292–306, 1976.
I. Prigogine. Time, structure, and fluctuations. Science, 201(4358):777–785,
1978.
Ilya Prigogine. From Being to Becoming Time and Complexity in the
Physical Sciences. W. H. Freeman, 1980.
Gunnar Pruessner. Self-Organised Criticality: Theory, Models and
Characterisation. Cambridge University Press, 2012.
D. Purves et al., editors. Neuroscience. Oxford University Press, 2018.
S. Rahmstorf. The thermohaline ocean circulation: A system with
dangerous thresholds? Climatic Change, 46(3):247–256, 2000.
K. Rauss, S. Schwartz, and G. Pourtois. Top-down effects on early visual
processing in humans: A predictive coding framework. Neuroscience &
Biobehavioral Reviews, 35(5):1237–1253, 2011.
E. Ravasz and A.-L. Barabási. Hierarchical organization in complex
networks. Physical Review E, 67(2):026112, 2003.
C. R. Reid et al. Army ants dynamically adjust living bridges in response to
a cost-benefit trade-off. Proceedings of the National Academy of
Sciences, 112(49):15113–15118, 2015.
D. Rind. Complexity and climate. Science, 284(5411):105–107, 1999.
Gerald H. Ristow. Pattern formation in granular materials. Springer, 2000.
James William Rohlf. Modern Physics from a to Z0. Wiley, 1994.
N. Rooney, K. McCann, G. Gellner, and J. C. Moore. Structural asymmetry
and the stability of diverse food webs. Nature, 442(7100):265, 2006.
A. Rosenblueth, N. Wiener, and J. Bigelow. Behavior, purpose and
teleology. Philosophy of Science, 10(1):18–24, 1943.
Don Ross. Philosophy of Economics. Palgrave Macmillan, 2014.
M. Rosvall and C. T. Bergstrom. Maps of random walks on complex
networks reveal community structure. Proceedings of the National
Academy of Sciences, 105(4):1118–1123, 2008.
G. Sargut and R. G. McGrath. Learning to live with complexity. Harvard
Business Review, 89(9):68–76, 2011.
M. Scheffer. Complex systems: Foreseeing tipping points. Nature, 467
(7314):411–412, 2010.
M. Scheffer, S. R Carpenter, V. Dakos, and E. H. van Nes. Generic
indicators of ecological resilience: Inferring the chance of a critical
transition. Annual Review of Ecology, Evolution, and Systematics,
46:145–167, 2015.
T. C. Schelling. Models of segregation. The American Economic Review, 59
(2):488–493, 1969.
E. Schneidman, M. J. Berry, R. Segev, and W. Bialek. Weak pairwise
correlations imply strongly correlated network states in a neural
population. Nature, 440(7087):1007–1012, 2006.
T. Schwander et al. Nature versus nurture in social insect caste
differentiation. Trends in ecology & evolution, 25(5):275–282, 2010.
Thomas D. Seeley. The Wisdom of the Hive: The Social Physiology of
Honey Bee Colonies. Harvard University Press, 2009.
Thomas D. Seeley. Honeybee Democracy. Princeton University Press,
2010.
C. R. Shalizi and J. P. Crutchfield. Computational mechanics: Pattern and
prediction, structure and simplicity. Journal of Statistical Physics,
104(3): 817–879, 2001.
C. R. Shalizi and K. Klinkner. An algorithm for building markov models
from time series, 2003. http://bactra.org/CSSR/.
C. R. Shalizi, K. L. Shalizi, and R. Haslinger. Quantifying self-organization
with optimal predictors. Physical Review Letters, 93(11):118701, 2004.
C. E. Shannon. A mathematical theory of communication. Technical report,
Bell Labs, 1948.
H. A. Simon. The architecture of complexity. Proceedings of the American
Philosophical Society, 106(6):467–482, 1962.
H. A. Simon. How complex are complex systems? Proceedings of the
Biennial Meeting of the Philosophy of Science Association, 2:507–522,
1976.
Adam Smith. An Inquiry into the Nature and Causes of the Wealth of
Nations. Strahan, 1776.
E. Smith. Emergent order in processes: The interplay of complexity,
robustness, correlation, and hierarchy in the biosphere. In C. H.
Lineweaver, editor, Complexity and the Arrow of Time. Cambridge
University Press, 2013.
Eric Smith and Harold J Morowitz. The Origin and Nature of Life on Earth:
The Emergence of the Fourth Geosphere. Cambridge University Press,
2016.
Kim Sneppen and Giovanni Zocchi. Physics in Molecular Biology.
Cambridge University Press, 2005.
D. Sornette. Predictability of catastrophic events: Material rupture,
earthquakes, turbulence, financial crashes, and human birth. Proceedings
of the National Academy of Sciences, 99:2522–2529, 2002.
Didier Sornette. Why Stockmarkets Crash: Critical Events in Financial
Markets. Princeton University Press, 2003.
Didier Sornette. Critical Phenomena in Natural Sciences: Chaos, Fractals,
Selforganization and Disorder: Concepts and Tools. Springer, 2nd
edition, 2009.
G. J. Stigler. The development of utility theory. II. Journal of Political
Economy, 58(5):373–396, 1950a.
G. J. Stigler. The development of utility theory. I. Journal of Political
Economy, 58(4):307–327, 1950b.
D. B. Stouffer and J. Bascompte. Compartmentalization increases food-web
persistence. Proceedings of the National Academy of Sciences, 108(9):
3648–3652, 2011.
M. Strevens. Complexity theory. In P. Humphreys, editor, The Oxford
Handbook of Philosophy of Science. Oxford University Press, 2016.
Steven H. Strogatz. Nonlinear Dynamics and Chaos: With Applications to
Physics, biology, chemistry, and engineering. Westview Press, 2014.
G. Sugihara and R. M. May. Applications of fractals in ecology. Trends in
Ecology & Evolution, 5(3):79–86, 1990.
T. Sullivan. Embracing complexity. Harvard Business Review, 89(9):88–92,
September 2011.
Nassim Nicholas Taleb. The Black Swan: The Impact of the Highly
Improbable. Random House, 2007.
A. Toomre and J. Toomre. Galactic bridges and tails. The Astrophysical
Journal, 178:623–666, 1972.
A. M. Turing. The chemical basis of morphogenesis. Philosophical
Transactions of the Royal Society of London. Series B, Biological
Sciences, 237 (641):37–72, 1952.
S. Ulam. Random processes and transformations. In Proceedings of the
International Congress on Mathematics, volume 2, pages 264–275,
1952.
P. A. M. Van Dongen. The Central Nervous System of Vertebrates. Springer,
1998.
Nicolaas Godfried Van Kampen. Stochastic Processes in Physics and
Chemistry, volume 1. Elsevier, 1992.
R. van Steveninck et al. Reproducibility and variability in neural spike
trains. Science, 275(5307):1805–1808, 1997.
A. J. Veraart et al. Recovery rates reflect distance to a tipping point in a
living system. Nature, 481(7381):357–359, 2012.
John von Neumann. Theory of Self-Reproducing Automata. University of
Illinois Press, 1966. edited and completed by Arthur W. Burks.
John von Neumann and Oskar Morgenstern. Theory of Games and
Economic Behavior. Princeton University Press, 1947.
Mitchell Waldrup. Complexity: The Emerging New Paradigm at the Edge of
Order and Chaos. Prentice Hall, 1992.
N. S. Ward. Functional reorganization of the cerebral motor system after
stroke. Current Opinion in Neurology, 17(6):725–730, 2004.
N. W. Watkins et al. 25 years of self-organized criticality: Concepts and
controversies. Space Science Reviews, 198(1-4):3–44, 2016.
M. L. Weitzman. On diversity. The Quarterly Journal of Economics,
107(2): 363–405, 1992.
G. Weng, U. S. Bhalla, and R. Iyengar. Complexity in biological signaling
systems. Science, 284(5411):92–96, 1999.
C. Werndl. What are the new implications of chaos for unpredictability?
The British Journal for the Philosophy of Science, 60(1):195–220, 2009.
ISSN 0007-0882.
B. T. Werner. Complexity in natural landform patterns. Science, 284(5411):
102–104, 1999.
G. B. West, J. H. Brown, and B. J. Enquist. A general model for the origin
of allometric scaling laws in biology. Science, 276(5309):122–126, 1997.
J. G. White, E. Southgate, J. N. Thomson, and S. Brenner. The structure of
the nervous system of the nematode Caenorhabditis elegans.
Philosophical Transactions of the Royal Society London B, Biological
Sciences, 314 (1165):1–340, 1986.
G. M. Whitesides and R. F. Ismagilov. Complexity in chemistry. Science,
284(5411):89–92, 1999.
Norbert Wiener. Cybernetics: Or Control and Communication in the Animal
and the Machine. MIT Press, 2nd edition, 1961.
K. Wiesner, M. Gu, E. Rieper, and V. Vedral. Information-theoretic lower
bound on energy cost of stochastic computation. Proc. R. Soc. A, 468
(2148):4058–4066, 2012.
J. Wilson. Metaphysical emergence: Weak and strong. In T. Bigaj and C.
Wuthrich, editors, Metaphysics in Contemporary Physics, pages 251–
306. Poznan Studies in the Philosophy of the Sciences and the
Humanities, 2015.
S. Wolfram, editor. Cellular Automata: Proceedings of an Interdisciplinary
Workshop, 1984. Center for Nonlinear Studies, Los Alamos, North-
Holland.
Stephen Wolfram. A New Kind of Science. Wolfram Media, 2002.
D. Wolpert. Information width: A way for the second law to increase
complexity. In C. H. Lineweaver, editor, Complexity and the Arrow of
Time. Cambridge University Press, 2013.
World Health Organization. Severe acute respiratory syndrome (SARS):
Status of the outbreak and lessons for the immediate future. Technical
report, 2013. www.who.int/csr/media/sars_wha.pdf.
Worldwidewebsize. Daily estimated size of the World Wide Web.
http://www.worldwidewebsize.com/, 2017. accessed 14 July 2017.
J. Z. Young. The number and sizes of nerve cells in octopus. Proceedings of
the Zoological Society of London, 140(2):229–254, 1963.
J. Ziv and A. Lempel. A universal algorithm for sequential data
compression. IEEE Transactions on Information Theory, 23(3):337–343,
1977.
G. Zöller, S. Hainzl, and J. Kurths. Observation of growing correlation
length as an indicator for critical point behavior prior to large
earthquakes. Journal of Geophysical Research: Solid Earth,
106(B2):2167–2175, 2001.
W. H. Zurek. Algorithmic randomness and physical entropy. Physical
Review A, 40(8):4731–4751, 1989.
Index
Adaptive behaviour, 2, 4, 19, 37, 43, 59, 61, 65, 66, 70, 72, 73, 82, 83, 85,
123, 124, 126, 129, 130
Agent, 5, 6, 27, 28, 45, 47, 49–52, 67, 72, 73, 94, 119
irrational, 48
rational, 47, 48
Agent-based models, 49, 94
Albert, Reka, 106
Algorithm, 41, 112, 116, 128
clustering, 107
community detection, 108
preferential attachment, 109
Algorithmic complexity, 110, 114–116, 135, 139
Algorithmic Information Theory, 139
Analogy between bee hives and brains, 44
Anderson, Philip, 3, 16, 22, 24, 81, 84
Ant colonies, 4, 38–41, 43, 81, 96, 126
Approximation, 5, 21–23, 116
Aristotle, 11
Attractors, 15, 103
Bak, Per, 123
Bank of England, 1, 52
Barabási, Albert-Lásló, 106
Beehives, 42–44, 69
compared to brains, 44
Beinhocker, Eric, 49
Belousov-Zhabotinsky (BZ) reaction, 4, 27, 70, 72–74, 77, 80, 128, 129
Bennett, Charles, 16, 80, 82, 115, 124
Bigelow, Julian, 12
Birds, 4, 6, 7, 67, 70, 71, 79, 82, 97, 125
Bouchaud, Jean-Philippe, 49
Boundaries, 23, 25, 32, 34, 64, 127
Brain, 6–8, 17, 26, 27, 44, 57–61, 66, 67, 69–71, 79, 81, 83, 85, 88, 91, 92,
98, 102, 105, 118, 119, 121, 124, 128, 129
Browder, Felix, 16
Buffon’s Needle, 41
Butterfly effect, 14
BZ reaction, see Belousov-Zhabotinsky reaction.
C. Elegans, 57
Causal interactions, 6, 67, 81, 84, 101, 115, 116
CDO, see Collateralised debt obligation.
CDS, see Credit default swap.
Cellular automata, 12, 15, 16, 113
Chaos, 14, 15, 64, 78, 100, 122
Chemistry, 2, 11, 12, 16, 19, 21, 27, 74, 77, 102, 118, 120, 127
Chomsky hierarchy, 112
Climate system, 33–37, 103
Collateralised debt obligation (CDO), 51
Communication, 6, 12, 38, 56, 67, 79, 83, 137
Complex adaptive behaviour, 2, see also Adaptive behaviour.
Complex networks, see Networks.
Complex Systems Society, 17
Complexity
conditions for, 10, 65, 66, 129, 130
functional conceptions of, 124
generic conceptions of, 121
physical conceptions of, 122
products of, 65
truisms of, 9, 12, 21, 28, 65, 74, 75, 88, 118, 126, 127
Complicated, 3, 7, 22, 33, 34, 63, 64, 78, 101, 117, 126
Computation, 5, 7, 9, 12, 13, 15–17, 76, 82, 83, 85, 88, 95, 111–113, 119–
122, 124–127, 137
Computational mechanics, 111–113
Conrad, Michael, 17
Consciousness, 8, 17, 37, 57, 60
Conway-Morris, Simon, 125
Correlation, 24, 26, 27, 67, 68, 71, 82, 89, 91, 92, 96–99, 101, 104, 108,
128, 138
length, 104
Pearson correlation coefficient, 97, 101
Correlation function, 96, 97, 104, 105, 109, 138
Coupled maps, 27
Covariance, 96, 97, 104, 136, 138
Cowan, George, 16, 17
Cowan, Jack, 16
Credit default swap (CDS), 51, 52
Critical phenomena, 3, 26–28, 67, 75, 127, 130, see also Self-organised
criticality
Critical point, 26, 27, 50, 104, 105, 123, 127, 128
Critical slowing down, 103–105
Crutchfield, James, 16, 111, 113
Cybernetics, 12, 13
Descartes, René, 19, 20
Differential equations, 14, 22, 63, 93, 96, 100, 102
Disorder, 10, 65, 68–70, 73, 85, 89–92, 95, 111, 122, 130
Distribution
degree distribution, 55, 56, 76, 91, 109, 141
frequency distribution, 141
Gaussian, 100
of wealth, 46
of words in English, 46, 138
power-law, 49, 58, 100
scale-free, 55, 56
Zipf, 100
Diversity, 30, 52, 53, 58, 65, 68, 70, 73, 85, 89, 92, 93
of type, 89, 92, 93
Dynamical systems theory, 8, 12–14, 16, 17, 100, 101, 103, 121
Dynamics, 7, 14, 15, 22, 23, 25, 27, 31–35, 49, 61, 70, 74, 75, 77–79, 89,
93, 94, 100–103, 110, 121, 123, 126–128
Economics, 2, 50
behavioural economics, 48
classical economics, 45
financial economics, 52, 53
macroeconomics, 50
microeconomics, 47, 50 164
supply and demand curves in, 45
Economies, 124
decision making economies, 44
digital economies, 47
Econophysics, 50
Ecosystems, 52, 92, 103, 105, 124
financial, 52
marine, 105
Effective complexity, 87, 114, 115
Effective measure complexity, 87, 98
Eigen, Manfred, 16
Emergence, 3, 4, 15, 16, 21–24, 27, 28, 37, 60, 65, 73, 74, 76, 78, 79, 81,
84, 85, 121–123, 127, 128
complexity as, 121
epistemological emergence, 121
of dynamics, properties and laws, 74
of structure at different scales, 74
of various kinds of invariance and forms of universal behaviour, 75
ontological emergence, 121
Energy, 21, 25, 27, 29, 36, 54, 59, 67–69, 72, 73, 99
free energy, 25, 71, 72, 123
Entropy, 33, 137
entropy power, 92
entropy rate, 91, 92, 111, 138
excess entropy, 98, 113
Shannon entropy, 90–92, 97, 113, 114, 137, 140
thermodynamic entropy, 71, 73, 123
Equilibrium, 19, 27, 28, 46–50, 52, 64, 73, 102, 123, 127
chemical, 73, 96
dynamic, 28, 70, 72, 73, 77
market, 46
Nash, 28
price, 45, 46
stable, 45, 53
static, 28
thermodynamic (thermal), 28, 65, 71–73, 121, 127
Erdös-Renyi random graph model, 90, 91, 107
Error correction, 43, 71, 79, 80
Eusocial insects, 37, 43, 83
Expectation value, 136, 137
Farmer, Doyne, 16, 17, 49, 50
Feedback, 4, 7, 8, 10, 12, 14, 29–31, 35, 36, 39, 40, 42, 43, 46, 47, 59–61,
65, 70, 71, 73, 79, 81, 83, 85, 93–95, 100, 101, 109, 110, 127, 129,
130, 133
negative, 35, 37, 43–45, 47, 71, 129
positive, 35–38, 46, 71, 128, 129
Feldman, Marcus, 16
Fixed point, 102, 103
Flocking/flocks, 70, 71, 97
Food webs, 55, 102, 105, 107, 140
Forces, 20, 29, 32, 34, 67, 118
Fractal, 16, 24, 25, 108, 109
Frauenfelder, Hans, 16
Function, 10, 66
Function (purpose), 2, 10, 19, 37, 58, 59, 63, 66, 71, 72, 79, 81–83, 85, 107,
124, 125, 127, 128, 130, 131
Gaia hypothesis, 130
Galileo, 20
Game of Life, 15, 121
Game theory, 17, 28
Gassendi, Pierre, 19
Geanakoplos, John, 50
Gell-Mann, Murray, 2, 16, 87, 114, 115, 120
General system theory, 13
Girvan, Michelle, 108
Golgi, Camillo, 57, 58
Gordon, Deborah, 40, 41
Grassberger, Peter, 16, 111
Gravity, 20, 23, 25, 29, 30
Haldane, Andrew, 52, 53
Halley’s comet, 20
Hierarchical predictive coding, 60, 61
Hierarchy, 81
in life sciences, 81
of complexity, 33, 124
of information processing, 60, 61
of neural processing, 60
of organisation/structure, 28, 56, 58, 81, 84, 118
History, 10, 11, 23, 29, 30, 32, 33, 66, 73, 81, 82, 88, 110, 111, 115, 116,
123, 125, 128, 130
evolutionary, 42, 57, 58, 82
of complexity science, 11, 12, 21, 63, 73
Holland, John, 17, 82
Holldobler, Bert, 37
Honeybees, 41, 42, 57, 66, 96
Huberman, Bernardo, 17
Ideal gas, 22, see also Law.
Idealisation, 5
Immune system, 1, 17, 83, 85, 118, 124
Information processing, 6, 9, 40, 42, 44, 58–61, 65, 75–77, 88, 119, 126
Information theory, 8, 76, 77, 82, 89, 90, 97, 130, 135
Inhomogeneity, 24, 69, 89, 127
Invisible hand, 45
Jeong, Hawoong, 106
Kahnemann, Daniel, 48
Kauffman, Stuart, 16, 17, 125
Kertesz, Janos, 49
Kleiber, Max, 99
Kondor, Imre, 49
Langton, Christopher, 16, 17
Law, 3, 6, 7, 13, 14, 20–23, 67, 75, 77, 79, 103, 118, 121, 126, 128
Boyle’s law, 3
emergence of, 74, 75, 80
Hubble’s law, 30
ideal gas laws, 3, 5, 22, 28, 74, 84, 127
Kepler’s laws, 22, 75
Newton’s laws, 13, 14, 20, 22
of co-existence, 22
of gravitation, 5, 20
of large numbers, 43
of supply and demand, 45, 46
of the pendulum, 5, 20
Pareto’s law, 46, 55
power law, see Distribution
second law of thermodynamics, 123
Life, 2, 8, 19, 25, 33, 76, 81–83, 116, 120, 121, 125, 127, 128, 130
Linearity, 50, 77, 78
Lloyd, Seth, 110, 111, 114, 115, 124
Log-log plot, 100
Logical depth, 82, 87, 140
Logistic map, 14, 100, 113
Lorenz, Edward, 14
Los Alamos National Laboratory, 16, 17
Lotka-Volterra model, 93, 102
MacKay, Robert, 78
Macrostate, 71
Mainzer, Klaus, 77
Mandelbrot, Benoît, 49, 108
Mantegna, Rosario, 49
Markets, 2, 6, 44, 45, 48–50, 61, 70, 71, 83, 124
Markov chain, 95, 136
Markov model, 112, 136, 137
Matrix, 108
adjacency, 140
Jacobian, 102
stochastic, 95, 112
transition, 112
Maxwell’s equations, 22
Maxwell’s theory of electromagnetism, 20
Maynard Smith, John, 17
Measures
of complexity, 87, 88, 123
of correlation, 96, 97, 138
of disorder, 89, 91
of diversity, 92
of feedback, 93
of history, 82, 110
of memory, 110
of modularity, 107
of nested structure, 107
of non-equilibrium, 95
of nonlinearity, 99
of numerosity, 88
of order, 89, 91, 96
of predictability, 138
of randomness, 139
of robustness, 101
Memory, 10, 66, 73, 81–83, 85, 110, 113, 125, 130
in the brain, 59
of an ant colony, 40
Microstate, 71
Milgram, Stanley, 106
Mitchell, Melanie, 84, 85
Modularity, 1, 10, 52, 53, 66, 68, 73, 81, 83, 107, 108, 126, 130
‘More is different’, 3, 9, 25, 60, 88, 128, 129
Morgenstern, Oskar, 47
Mutual information, 97, 98, 101, 138
Nested structure, 10, 66, 68, 73, 81, 83, 107–110, 124, 128, 130
Networks, 6, 9, 54–57, 73, 75, 76, 78, 83, 85, 88, 106, 107, 109, 119, 126,
137, 140
average path length, 90, 91, 106
clusters in, 107
collaboration, 55
financial, 53, 106
genetic regulatory, 102
IT, 71, 124
metabolic, 55
neural, 21, 106
protein, 55, 106, 107
scale-free, 55
social, 5, 106, 107
transportation, 124
Neurons, 6, 8, 44, 57–60, 66, 70, 79, 82, 83, 92, 98, 105
Newman, Mark, 108
Newton, Isaac, 5
Nihilism about complex systems, 117–119
Non-equilibrium, 10, 50, 65, 71, 85, 95, 96, 122, 130
Nonlinear dynamical systems, 100
Nonlinearity, 10, 14, 15, 46, 64, 65, 73, 77–79, 83, 85, 93, 99–101, 105,
108, 109, 128–130
Numerosity, 10, 61, 65–68, 73, 85, 88, 126, 130
Orbits of Mercury, 20
Order, 4, 8–10, 22–29, 31, 45, 57, 65, 68–74, 76, 77, 79–82, 84, 85, 88, 93,
96–98, 101, 111, 113–116, 119, 121–123, 125–128, 130
Packard, Norman, 17
Page, Scott, 93
Pagels, Heinz, 111
Particles, 3, 5, 19–22, 26, 29, 30, 32, 66–68, 74, 75, 79, 118, 120
Perelsen, Alan, 17
Phase transitions, 3, 16, 19, 24–29, 50, 67, 68, 72, 75, 101, 103, 104, 107,
127, 128
Physics, 2–4, 11, 12, 16, 17, 19–25, 29, 31, 33, 49, 50, 61, 70, 74, 75, 84,
100, 103, 118, 121, 122, 127, 128
astrophysics, 3, 8, 33, 120
condensed matter physics, 22, 25, 28, 66, 71, 72, 76, 79, 85, 97, 123,
126, 127
statistical physics, 8, 21, 28, 125
Pines, David, 16
Poincaré, Henri, 14
Population growth, 14
Power law, see Distribution. Pragmatism about complexity science, 119
Prediction, 7, 10, 11, 13, 14, 20, 35, 37, 60, 61, 66, 98
Probability distribution, 89–92, 95, 99, 100, 105, 108, 112, 113, 135, 137
joint, 91, 98, 101, 108, 135, 138
Probability theory, 73, 89, 126, 135
Puck, Ted, 16
Purpose, see Function (purpose)
Quantum
chemistry, 3, 23, 116
classical limit, 67
computation, 124
dynamics, 78
fields, 20, 22, 127
gravity, 21
mechanics, 22
multiverses, 124
numbers, 22
theory, 21
Quorum decision making, 41, 43, 44, 61
Radiation, 2, 5, 19–21, 28–31, 34, 36, 46, 50, 72, 104
Ramón y Cajal, Santiago, 57
Ramsey, Norman, 16
Randomness, 59, 89, 114, 124, 137, 139, 140
Reaction-diffusion model, 73
Realism about complex systems, 120, 125
Reductionism, 77, 84
Relativity, 21
Resilience, 79, 101, 103, see also Robustness.
Rich-get-richer effect, 46, 109
Robustness, 1, 10, 56, 66, 70, 73, 79, 80, 83, 85, 101, 106, 107, 110, 124–
130, see also Resilience.
functional, 59, 83, 101
of algorithm, 101
structural, 101, 107
Rosenblueth, Arturo, 12, 13
Rota, Gian-Carlo, 16
Sandpile model, 27
Santa Fe Institute, 17, 73
Scale
characteristic scale, 68
length scale, 24, 28, 29, 31, 51, 68, 75, 77, 104, 114, 128
time scale, 2, 3, 10, 21, 24, 28, 30–33, 35–37, 46, 49, 59, 65, 68, 70, 79,
80, 82, 93, 94, 103, 104, 127, 128
Scale invariance, 24, 26, 105
Schuster, Peter, 17
Scientific Revolution, 11
Scott, Alwyn, 16
Self-organisation, 4, 12, 16, 19, 27, 29, 37, 76–79, 84, 96, 113, 123, 127,
129, 130
Self-organised criticality, 27, 50, 105, 106, 123, 128
Self-similarity, 123
Sensitive dependence on initial conditions, 14
Shannon entropy, see Entropy.
Shannon, Claude, 137
Simon, Herbert, 81, 115, 124
Singer, Jerome, 16
Six degrees of separation, 106
Small-world effect, 55
Smith, Adam, 45
Smith, Vernon, 48
Snowflakes, 23
Stability analysis, 102, 103
Stanley, Eugene, 49
Statistical complexity, 87, 110
Statistical mechanics, 22, 24, 26, 72, 76, 97, 103, 111, 114, 121, 126
Sugarscape model, 94
Superposition principle, 14
Symmetry, 24, 69, 77, 81
Symmetry breaking, 24, 32, 77, 127
Thermodynamic depth, 124
Thermodynamics, 24, 25, 72, 95, 122, 123, see also Law.
Tipping points, 103, 106
Toffoli, Tommaso, 16
True measure complexity, 111, 115
Truisms, see Complexity.
Turing machine, 119
Turing pattern, 73
Turing, Alan, 73, 119, 139, 140
Tversky, Amos, 48
Ulam, Stanislaw, 15
Universal behaviour, 6, 9, 29, 75, 85, 88, 119, 126
Universe, 1, 19, 23, 28–33, 76, 82, 85, 116, 121, 123
Utility function, 47
Variance, 89–93, 97, 100, 136
von Bertalanffy, Ludwig, 13
von Neumann, John, 15, 47
Waggle dance, 42, 43
Water, 2, 7, 21, 23–26, 28, 29, 33–35, 42, 43, 68, 74, 80
Weather, 2, 14, 32–34, 36, 68, 127, 129
Wiener, Norbert, 12, 13
Wilczek, Frank, 16
Wilson, Edward, 37
Wolfram, Stephen, 16
World Wide Web, 46, 53–56, 90, 106, 109
Young, Karl, 111
Zhang, Yi-Cheng, 49 169
from
Your gateway to knowledge and culture. Accessible for everyone.
z-library.se singlelogin.re go-to-zlibrary.se single-login.ru
O cial Telegram channel
Z-Access
https://wikipedia.org/wiki/Z-Library
ffi