Introduction To Systems Analysis Mathematically Modeling Natural Systems
Introduction To Systems Analysis Mathematically Modeling Natural Systems
•
Dieter M. Imboden • Stefan Pfenninger
123
Dieter M. Imboden Stefan Pfenninger
Department of Environmental Systems International Institute for Applied Systems
Science Analysis (IIASA)
ETH Zurich Laxenburg
Zurich Austria
Switzerland
v
vi PREFACE
1 Introduction 1
1.1 Systems Analysis . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 What is a System? . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 What is a Model? . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 The Formation of Models in Natural Science . . . . . . . . . 9
1.5 Questions and Problems . . . . . . . . . . . . . . . . . . . . 13
3 Static Models 33
3.1 Equilibrium Distribution Between Water and Air . . . . . . 34
3.2 Equilibrium Distribution Between Water and Sediment . . . 35
3.3 Multi-dimensional Static Models . . . . . . . . . . . . . . . 37
3.4 Questions and Problems . . . . . . . . . . . . . . . . . . . . 39
vii
viii CONTENTS
C Formulary 233
C.1 Linear Inhomogeneous First-Order Differential
Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
C.2 System of Two Linear First-Order Differential
Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
C.3 General Solution of the Linear Second-Order
Differential Equation with Constant Coefficients . . . . . . 235
C.4 Solution of Linear Differential Equations
with Imaginary Eigenvalues . . . . . . . . . . . . . . . . . . 237
D Eigenvalues 239
D.1 The n-Dimensional System . . . . . . . . . . . . . . . . . . 239
D.2 Explicit Solution for the Two-Dimensional System . . . . . 240
Bibliography 247
Index 249
Chapter 1
Introduction
Dang
and matrix calculus, linear algebra, and even take a glimpse at nonlinear
and partial differential equations. Most of the mathematical methods used
are covered in the appendices. Their treatment there is brief however, and
without proofs. Therefore it will not replace a good mathematics textbook
for someone who has not encountered this level of math before.
Although most of the examples used here are drawn from the environ-
mental sciences, this book is not an introduction to the theory of aquatic
or terrestrial environmental systems. Rather, a key goal of the book is to
demonstrate the virtually limitless practical potential of the methods pre-
sented. Readers are encouraged to use the tools provided to tackle problems
from their own areas of interest.
The first chapter is a short overview of the philosophy behind building
models in science and does not use mathematical terms yet.
Book website
Visit the book website at www.systems-analysis.org. You’ll find interactive versions
of several examples to help you understand the behavior of systems. You can also find
additional resources there, such as a list of further recommended books.
mathematical methods.
The principle underlying all of this is called reductionism. The reduc-
tionistic principle is not without its critics, and sometimes it is given re-
sponsibility for spawning a sort of technological progress which is blind to
its negative consequences (environmental impacts such as biodiversity loss,
climate change, as well as other effects such as the growing division between
rich and poor).
We do not claim to be able to contribute to this important discussion.
We do argue, however, that to criticize, improve or even transcend an idea,
one must first understand it. On closer inspection, we discover that charac-
terizing a scientific approach as reductionistic or on the contrary as holistic
is always a question of one’s point of view. A cell biologist would probably
call a molecular biologist’s work reductionistic. Cell biology, however, will
appear as a holistic field to her. On the other hand, from a population
ecologist’s viewpoint, cell biology would itself appear reductionistic.
The division of science into different areas of expertise thus not only
takes place through by putting different disciplines into different drawers of
a chest (see illustration in the margin), but each drawer is again subdivided
into further drawers. The chest of drawers depicted is therefore only one
layer of reductionistic subdivision. In reality, many such chests would have
to be drawn and each of their drawers would be subdivided further and
further. Of course, this nested mass of drawers is not only a feature of
biology—the reductionistic approach is used in every scientific field.
Even though it is not our aim to highlight this debate in principle, it will
play a role indirectly throughout the book. The debate around reductionism Reductionism can be
forces us, time and again, to carefully examine the methods we develop thought of as a chest di-
here. As yet there are few concepts with which we could comprehend the vided into drawers that are
world in a holistic way (whatever that may mean), without falling back to themselves further subdi-
the method of subdividing it into smaller and more understandable parts. vided.
Even if we should develop such methods in the future, analyzing a system
by dissecting its individual parts will always remain an important part
of science. In any case, we want to think critically about our use of a
reductionistic approach, and not lose track of the bigger picture.
In this book we want to convey how to model and analyze real-world
systems. The underlying core principle is to construct simplified represen-
tations of such complex systems. These representations, called models,can
often be expressed mathematically (e.g. through a system of equations).
The construction of a mathematical model can be reduced to a few core
ideas, in particular:
With this book you, the reader, should gain the ability to analyze real
systems and make quantitative statements about their behavior. This re-
quires the ability to translate empirically identified properties of these sys-
tems into a mathematical form, and to discuss the resulting equations (in
particular, differential equations), in some cases even to solve them an-
alytically. We will also delve into the behavior of more complex models,
without specifically discussing how to solve them. Finally, we also want to
demonstrate the limits of simulation and the emergence of mathematical
artifacts.
Before we look at different mathematical models in more detail, we first
have to clarify what exactly a system and a model are, and to understand
the significance of modeling in science.
A system is a set of objects The word system originates from Greek and means configuration or compo-
between which relations sition. The New Oxford American Dictionary gives the following definition:
exist. “a set of connected things or parts forming a complex whole”. In a more
abstract way, one could also say that a system is a set of objects between
which relations exist.
A system is more than the What does this mean? A system is composed of different parts, the
sum of its parts. objects or system components. They are connected with each other through
mutual interactions which we call internal relations.
A pile of metal parts is therefore not yet a system. The parts have to be
connected in a meaningful manner. They might be connected to form the
system clockwork. If one cog is removed from that system, it will entirely
cease to function. This fact is reflected in the saying “a system is more than
the sum of its parts”.
WHAT IS A SYSTEM? 5
Only through internal relations can the constituent parts become a sys-
tem. This is not always as straightforward as in the clockwork example:
But setting the system boundary is not always as simple, as the next ex-
ample demonstrates:
Summing up, we can say that systems are theoretical constructs that help
us in understanding a part of the world. There is no absolutely valid way to
define a system’s boundary. In fact, the exact choice of boundary depends
on the question we are asking.
WHAT IS A MODEL? 7
We can trace the origin of our word model to ancient Rome, where Modulus is Latin for
small replicas of buildings were called modulus. Just as Roman architects “small-scale replica of a
needed replicas to assist them in designing large and complex buildings, if building”.
we want to analyze a natural system, we must first describe it in simplified
terms. Whether this is as a drawing, a functional diagram, a mathematical
equation or a verbal account, we can call such a description a model. So
we can say that a model is always a simplified image of a real system.
A model, therefore, does not necessarily have anything to do with maths
at the outset. It is simply a concept with which we can describe a com-
plex system in simplified terms. Models are used in architecture to plan
buildings, or in chemistry to describe atoms as spheres and the molecular
bonds as little rods between them (Fig. 1.2). Models are in use everywhere,
whether they are physical models or purely mental constructs.
A model is the lens through which we look at the real system—it is not
an exact copy of that system. For the same system, there can be completely
different models. We could depict a landscape as a topographical map, or as
an impressionistic painting. In both cases, the simplification of the system
“landscape” is our explicit intention. It is impossible to say a priori which
model is better. One model is not superior to another because it is more
complex. Depending on the point of view that we want to emphasise and the
question we intend to answer, we include certain properties of the system
in the model while excluding others. You can think of it like a dog taken for
a walk in a forest. The dog’s model of the forest will be different than that
of its owner; it might predominantly consist of scents, for instance. But he
will be able to navigate the forest just as well or even better than his owner.
Good models allow us, through certain simplifications, to arrive at a
readily comprehensible representation of the real world that can be de-
scribed mathematically. Such models have a long tradition in science and
engineering. The only way to apprehend and analyze a real system is with
a model, that is, by framing it in terms that hold relevance for us. Often,
an important motivation to construct a model is to make predictions about
a system’s future behavior. As we will see, however, making predictions is
not the only role that models play.
Using this example, we can identify the two main tasks of systems analysis:
Model formation: The construction of a model begins with laying
down the system boundary, the definition of the system variables, and the
internal and external relations. The starting point is often empirical data,
for instance from well-designed experiments. These data are analyzed in
view of constructing a model. The model, usually a mathematical one,
should explain the data in the simplest possible form, in terms of the in-
ternal model structure and the external relations. In other words: systems
analysis looks for the (simplest) structure of the “black box” that can re-
produce the postulated connection between the state of the environment
and the system’s state (Fig. 1.3).
Fig. 1.3: As a first step in mathematical model formation, the external relations
(state of the environment) and the system variables (Vi ) have to be defined and
quantified. The black box model attempts to describe the system variables from
the external state as simply and meaningfully as possible
MODEL FORMATION 11
In the end, therefore, gathering data only makes sense if further insight
results from it, either as a generalizable statement, a prognosis, or as the
formulation of new questions, that is, the planning of new experiments. To
process observations in this manner, we need models.
Box model
I ve built a 3-dimensional box model. For the most dramatic demonstration, Wow! Wait a second, where is my money?
It conveys a deeper understanding of we need a modest input of money
how money flows in our society. similar to how we pay taxes.
Fascinating,
isn t it?
Observations
Simulation Measurements
12 INTRODUCTION
With this, we’ve reached the end of the first chapter, in which we looked
at the importance of the modeling method in science. In the next chapter,
we will get to know the fundamental tools with which we can actually build
models.
At the end of each chapter, there is a set of questions and problems. In
the questions, we recap the chapter’s contents. They can usually be quickly
and easily answered. The problems however require more in-depth work,
but will greatly help with retaining and applying the key concepts and tools
introduced in the chapter.
QUESTIONS AND PROBLEMS 13
Question 1.2: What is implied with the statement “The relation between
a system and its environment is asymmetrical”?
Question 1.3: Give a short definition for each of the following terms:
system, model, mathematical model.
Question 1.5: Give some of the most important decisions that have to be
made when constructing a model.
The dimension of a model The number of system variables is called the dimension of a model. The
equals the number of sys- example above represents a three-dimensional model with the three system
tem variables. variables Maq , Msed and Mbio . In order to formulate the model mathemat-
ically, we must describe the internal relations (the mass flows between the
three system variables) and the external relations. By doing so, we arrive at
the system or model equations. They usually contain additional quantities:
the model parameters. In the next section, we will illustrate the method to
create the system equations step by step with an example.
Phosphorus
Fig. 2.2: Static lake
input
model: Average phospho-
rus concentration in the
lake as a function of the
current phosphorus input
fluctuate with time. In a first attempt, we assume that there is an immediate Is there a correlation
relation between the phosphorus input at time t and the average phosphorus between Jin (t) and Caq (t)?
concentration in the lake measured at the same time. Mathematically, this
can be formulated as follows1 :
Caq (t) = Function of Jin = f Jin (t) (2.1)
We assume that the lake in question has been under surveillance for
several years and that a data set exists from which we can calculate Caq (t)
and Jin (t) for different times. To find the unknown function in Eq. (2.1), we
can sketch the values of Caq (t) and Jin (t) in a two-dimensional diagram.
Let’s look at a numerical example:
30
20
10
J in
0
0 1000 2000 3000 4000
Phosphorus input [ kg yr −1]
The above phosphorus model is called static, because for every value
of the external relation R (in this example, the input Jin (t)) there is a
precisely defined value of the system variable V. Mathematically speaking,
a static model with one system variable has the following form2 :
2Not in all cases can we explicitly describe the system variable V (as we did in
Eq. 2.4). It can also be given through an implicit equation, such as eaV = bV.
A FIRST LOOK 19
Equation (2.4) is called the system equation of the model. The function
f (R) can have an arbitrarily complicated form and can be defined by one
or several model parameters p. If the function f (R) is linear, as is the case
in Eq. (2.2), the model is called linear.
We have described the system “phosphorus in a lake” with a linear,
static model. By using Eq. (2.2), we can predict how the lake’s phospho-
rus concentration changes in response to a changed phosphorus input. We A static model cannot ad-
cannot, however, make any statement about how long it takes until the equately describe changes
phosphorus concentration reaches its new value after the input variable has through time.
been changed. In a static model, the change takes place instantaneously. A
dynamic model is needed in order to describe the transition from one state
to another.
Before we delve deeper into dynamic models in Sect. 2.3, we will briefly
discuss the importance of units and dimensions.
mg m−3
p = 1.33 × 10−2
kg year−1
mg m−3
= 1.33 × 10−8
mg year−1
= 1.33 × 10−8 year m−3
The parameter p therefore has the dimension of an inverse volume flow rate
(T L−3 ). We will get to know the physical significance of this parameter in
Sect. 4.2. For now, let’s look at another example:
20 MATHEMATICAL MODELS
103 g
0.01 kg m−3 = 0.01 × = 0.01 g L−1 = 10 mg L−1 (2.6)
103 L
Thus 10 mg of the toxic substance are dissolved in 1 L of water.
Most common spreadsheet and statistics software allows the user to graphi-
cally display the empirical relation between data pairs (like in Fig. 2.3) and
to estimate the function by a regression. If we relate data pairs of two or
more measured variables by a linear or nonlinear regression, a static model
is produced where the regression equation becomes the model equation.
Since setting up model equations by regression analysis is very easy,
one will often forget to verify whether the equation makes any physical
sense. Even in literature, one occasionally finds model equations that have
different dimensions on both sides. Furthermore, equations are often set up
in a way that makes them valid only for a specific set of units.
In order to demonstrate this point, let’s look at experimental data on
the growth of an algae population under varying nitrate concentrations in
water. The measurements were analyzed by a nonlinear equation.
A FIRST LOOK 21
For a serious modeler there are three ways to prevent this kind of confusion:
(1) The easiest, but not always the best way: We simply write the correct
units in square brackets behind each variable:
G (h−1 ) = 0.1 (C (mgN L−1 ))0.6 (2.8)
This notation makes the equation confusing and difficult to read.
(2) A better way is to replace the numerical factor by the symbol p and
to separately give the value of p together with the correct unit:
G = p C 0.6 , with p = 0.1 h−1 (mgN L−1 )−0.6 (2.9)
G (h−1 ) Specific growth rate
C (mgN L−1 ) Nitrate concentration
If we would now like to insert the concentration in mol m−3 , we have
to keep in mind that 1 mgN L−1 = 1 gN m−3 = 14 1
mol m−3 . Thus,
the parameter can be converted as follows:
p = 0.1 h−1 (mgN L−1 )−0.6
1
= 0.1 h−1 ( mol m−3 )−0.6
14
= 0.1 h−1 (14)0.6 (mol m−3 )−0.6
= 0.49 h−1 (mol m−3 )−0.6
22 MATHEMATICAL MODELS
Tabloid units
Editor: “You want a front page headline because of a laughable 10
Curies in RadonCorp’s waste?”—Journalist: “How about 400 billion Bec-
querel?”a —“Much better. You’ve got journalism in your blood after all!”
a Curie and Becquerel are both units for the radioactivity of a radiation source.
1 Becquerel corresponds to one decay per second. 1 Curie = 37 billion Bec-
querel.
Summing up, verifying the dimensions and units when we set up and
solve system equations is one of the easiest and most efficient ways to
prevent mistakes. The following rules may help:
• The algebraic expressions to the left and right of the equals sign must
always have the same dimension.
• Addition and subtraction are only possible if all algebraic terms have
the same dimension.
• When inserting concrete numbers into an algebraic equation, their
units have to be compatible. In other words, they have to be chosen
so that the same combination of units results on both sides of the
equals sign.
• Exponents in equations are non-dimensional (and therefore also with-
out units). The same holds for arguments in transcendental functions
A FIRST LOOK 23
such as sin(..), cos(..) or exp(..). Ensure that you only have fractional
powers or transcendental functions of non-dimensional variables in
your equations.
A dynamic model describes In general terms, a dynamic model can describe how rapidly the system
the system’s time- variable V changes when the external relation R varies. The dynamic model
dependent answer to an describes the time-dependent answer to the external change. Figure 2.4
external change. shows the adjustment of a linear system to a disturbance that occurs sud-
denly at time t = 0. The solid line represents the simulation with a dynamic
model. The system variable V(t) does not react immediately to the distur-
bance, but adjusts gradually to a new system state in equilibrium with
the changed external relation R. The static model, depicted as a dashed
line, disregards this adjustment behaviour; the system variable immediately
changes to the new system state.
If the temporal change of the system variable V is smooth, i.e. not made
up by sudden jumps,3 it can be described by the first derivative of the
system variable with respect to time. The mathematical model is therefore
4 In a first order (n-th order) differential equation, the function that consti-
tutes its solution is no higher than in its first (n-th) derivative. The models
we deal with here are usually first order with respect to time. In physics, par-
ticularly in classical mechanics, they are often second order with respect to
time. For the solution of these differential equations this distinction is only
marginally relevant, because one can convert an n-order differential equation
into n first order equations.
26 MATHEMATICAL MODELS
Ok, I basically
hate dogs.
This means that the state V(k+1) is calculated from V(k) according to
a specific “recipe”. The superscript index k in parentheses indicates the
system state at time tk . Unlike in a differential equation Eq. (2.12), the
system variable’s change in time is not described infinitesimally
by the first
derivative, but by the change occurring during the time step tk+1 − tk .5
A more detailed description of these models follows in Chap. 7.
With nonlinear differential Thus, apparently identical systems may develop differently. This is called
equations, almost identical deterministic chaos. The weather is a typical example for chaotic behavior.
initial conditions can later We are capable of predicting the weather over a period of a few days rela-
lead to completely different tively well. But our limited knowledge of the atmosphere’s state and other
system states. relevant factors means that a prognosis over 2 months will likely remain
impossible.
When modeling, we can use various techniques to give essentially
deterministic equations a whiff of stochastics. The simplest possibility is
to introduce a randomly varying external relation. In this book, however,
we will focus primarily on deterministic systems.
In this chapter, we got a general idea of some different model types and
thereby encountered some properties with which we can characterize mod-
els. These properties include the pairs static-dynamic, discrete-continuous
(either in space or in time) and deterministic-stochastic. In the following
chapters, we will discuss the different model types in more detail.
Deterministic Chaos
QUESTIONS AND PROBLEMS 31
Question 2.4: What does the dimension of a model mean? (Note: Com-
paring questions 2.3 and 2.4 will make clear that dimension has a different
meaning in mathematics than it does in physics.)
Question 2.6: Look for some examples of systems that contain a stochastic
component.
Question 2.7: Grubs live in the ground for 3 years and hatch as May
beetles in spring of the fourth year. Sketch a simple regional model for
May beetles. In particular, you should think about the following issues:
choice of the system variables, type of model (dynamic or static, in space
or in time, discrete or continuous).
(a) A factory is built on the lake shore. From time t0 onwards, it continu-
ously discharges a substance into the lake. This substance is volatile,
i.e. it can evaporate into the atmosphere, but is not degraded in the
lake. It doesn’t deposit in the sediment.
In both cases, you can assume that the substance’s concentration in the
inlet is zero.
32 MATHEMATICAL MODELS
σp [a−1 ]
.1
σp = 10
z
.01
1 10 100 1000
average depth z [m]
∂C(z, t) ∂2C ∂C
= −k1 + k2 2 + k3 + k4 C(C ∗ − C)
∂t ∂z ∂z
Determine the dimensions (expressed by (M,L,T)) of the parameters k1 ,
k2 , k3 , k4 and C ∗ . C is the concentration of a substance, z a length and
t the time. Hint: How is the dimension of a first or second derivative of a
variable determined?
Static Models
Static accounts
Anne to Bill: “Yesterday, I brought 10 dollars to the bank, and the teller
told me that I now had $100 in my account.”—Bill: “Hmm. . . So, if I bring
him 20 bucks tomorrow, my balance will be $200!”
In this chapter we want to take a closer look at static models. In Chap. 2 A static
we’ve seen that with a one-dimensional static model, the system variable V system?
has an exactly defined value for a given external relation R. A static model
therefore simply describes the equilibrium state between R and V and does
not give any indication about how rapidly V adjusts to changes in R. Thus,
static models are particularly useful when V quickly reacts to changes in R
(that is, in cases where we can neglect the time that V takes to change).
The data pairs are plotted in Fig. 3.1. They lie approximately on a
straight line, so we can formulate the following linear static model:
0.01
0.005
0
0 0.02 0.04 0.06 0.08 0.1 0.12
Benzene concentration Caq [g L−1]
EQUILIBRIUM DISTRIBUTION 35
William Henry
(1775-1836)
tribution coefficient is
Kd = 1.7 ± 0.008 L kg−1sed 0.15
0.1
0.05
0
0 0.02 0.04 0.06 0.08 1.1 0.12
Benzene concentration Caq [g L−1]
This experiment can also be repeated with other chemicals and particles.
The two-dimensional diagram representing the equilibrium distribution of
a chemical substance between water and a solid phase is called sorption
isotherm. In our example, the sorption isotherm is linear. The parameter
p is the slope of the sorption isotherm and is called distribution coefficient
Kd , with a dimension of [ L3 M−1 ]:
Csed
Kd = (3.4)
Caq
MULTI-DIMENSIONAL STATIC MODELS 37
Cair → R
Caq → V1
Csed → V2
To draw up the system equations, we use the relations derived in the two
previous examples. This yields:
1
V1 = R (3.5)
Ka/w
V2 = Kd V1 (3.6)
1
→ p1
Ka/w
K d → p2
Because we know R and the two system parameters, we can calculate the
benzene concentrations V1 and V2 directly with Eqs. (3.5) and (3.6). This
gives us:
1 1
Caq ≡ V1 = R= × 0.01 g L−1 = 0.05 g L−1
Ka/w 0.2
Csed ≡ V2 = Kd V1 = 1.7 L kgsed −1 × 0.05 g L−1 = 0.085 g kgsed −1
Where:
Question 3.2: Which of the following laws and formulas from physics,
chemistry and biology correspond to a static model?
Don’t worry if you do not know some or all of these laws or can’t
clearly classify them. They are not necessary to understand the material.
In addition, some of them will be discussed in detail later in the book.
Caq [μmol L−1 ] 0.17 0.51 1.8 3.6 7.6 19.5 26.5
1,4-Dinitrobenzene Cmin [μmol kg−1
min ] 241 633 1,640 2,850 4,240 6,100 7,060
Draw a two-dimensional diagram with the data pairs and try to find a
model with a maximum of two free parameters to describe the sorption.
Supply curve
Price per good
Demand curve
2. Supply for the good increases with increasing price (Fig. 3.3b)
(a) Use the two curves to graphically determine the resulting price.
(b) What happens to the price of the good if, for a constant
supply curve (Fig. 3.3b), the demand grows, i.e. the demand
curve (Fig. 3.3a) moves to the right?
(c) What happens to the price and the number of goods sold if the
manufacturing costs decrease (for a given price, more goods can
be manufactured)?
Chapter 4
In the remaining part of the book, we will exclusively deal with dynamic
models. As we already noted in the introduction, one of the principal char-
acteristics of natural systems is that they are not static.
Dynamic and continuous-time systems are usually described with differ-
ential equations. Although natural systems are hardly ever linear in reality,
we will first take a close look at linear models (and thus linear differen-
tial equations). The field of systems analysis is rooted in physics and the
technically-oriented sciences. In those fields, linear models play a particu-
larly important role. They have a key feature: they have analytical solutions
which can be linearly combined at will to construct new solutions. These
simple solutions are well-suited as building blocks for more complex models.
Real systems are usually nonlinear. Often, however, they can be approx- Nonlinear systems can of-
imated by linear models in a piecemeal way, that is, within limited ranges ten be approximated by lin-
of the system variable. As we will see in Chap. 6, the concepts developed for ear models piece by piece.
linear models serve as the point of departure for the analysis of nonlinear
models.
First, let us examine continuous-time systems with only one system
variable. The dynamic models resulting from this describe the change of a
single variable of the system. We do not spatially differentiate the system,
we treat it as one box, for instance as a thoroughly mixed volume. The
mathematical model for such a system is given by the following first-order
differential equation:
dV(t)
= R + pV(t) (4.1)
dt
Where:
V(t) System variable
R External relation
p Model parameter
dV
= {Production processes} − {Loss processes} (4.2)
dt
By “production” we mean all processes through which the value of the vari-
able V is increased. Apart from processes taking place within the system,
this also includes transport processes from the environment into the sys-
tem. “Loss” means all reaction and transport processes through which the
value of the variable V decreases.
If Eq. (4.2) describes a linear model, both process types consist of a sum
of zero-order and first-order processes:
{Production processes} = Jp + kp V
(4.3)
{Loss processes} = Jl + kl V
Jp and Jl are the zero-order rates independent of V, while (kp V) and (kl V)
are the first-order rates that are proportional to V. The parameters kp and
kl are called first-order specific (conversion) rates. They always have the
dimension [ T−1 ] independently of V.
Balance equation
Production processes and loss processes dv/dt is not looking good at all. We don t like it when our students
are not balanced in other words, we are know less at the end of the semester
losing knowledge. than at the beginning.
Ok.
I see.
Oops...
dV
= (Jp + kp V) − (Jl + kl V)
dt
= (Jp − Jl ) + (kp − kl ) V (4.4)
= J + k V
2 From now on, to make things easier to read, we will usually not write the
time dependence of the system variables. The notation V will stand for V(t).
46 LINEAR MODELS WITH ONE VARIABLE
The model consists of three production and loss terms. One of them is
zero-order:
J = Jp = 500 kg year−1
The two others are first-order, they have the specific net rate:
So, overall:
dM
= 500 kg year−1 − (0.2 year−1 ) M
dt
We will return to this in Example 4.3.
dV
= J + k V (4.5)
dt
It has the initial condition V(0) = V0 and the following solution (as long
as k is not zero):
J k t J
V(t) = (V0 + )e − for k = 0 (4.6)
k k
For the homogeneous differential equation
dV
= k V (4.7)
dt
the solution is accordingly:
V(t) = V0 ek t
(4.8)
By inserting Eq. (4.6) or (4.8) into the original differential equation, you
can easily validate the result given above. The initial condition V(t = 0) =
V0 is correctly expressed, as can be seen by setting t = 0 in the solutions.
Let’s first discuss the solution of the homogeneous Equation (4.8). As
can easily be seen, V(t) behaves entirely different for t −→ ∞ depending
on whether k is negative or positive (we will not discuss the trivial case
of k = 0). The two cases are shown in Fig. 4.1. They are the exponential
3 To be precise, this is only true for deterministic models (see Sect. 2.6).
48 LINEAR MODELS WITH ONE VARIABLE
growth and decay curves. For the case of a negative k we introduce the
new specific rate k (which we will use again later):
k = −k , k>0 (4.9)
To identify the exponential curve it helps to divide both sides of Eq. (4.8)
by V0 and then to calculate the natural logarithm on both sides:
V(t)
ln( ) = ln V(t) − ln V0 = ln(ek t ) = k t
V0
or:
ln V(t) = ln V0 + k t (4.10)
Thus, if we plot ln V(t) against time t, the curve should be a straight line
with slope k (see Fig. 4.1).
1
0
0 0.1
Relative system variable
0 −k t
(t) = e ln (t)
1
k = −k > 0
1
0.1
0 0.01
Time t →
Global population
The differential equation of the exponential growth model is analogous (billion people)
to Eq. (4.7):
dN
= kp · N (4.11)
dt
We arrive at the solution with Eq. (4.8):
Solving Eq. (4.12) for kp and using t = 125 year, N (t)/N 0 = 4, we get:
1 N (t) 1
kp = ln = ln 4 ≈ 0.01 year−1 (4.13)
t N0 125 year
In many models, the specific total rate k is negative. With the definition
Eq. (4.9) we can write the resulting differential equation in the following
form, also replacing J with J to simplify further:
dV
=J −k V , k>0 (4.14)
dt
This equation has the solution:
J −kt J
V(t) = (V0 − )e + , k>0 (4.15)
k k
Cheeeeese!
steady state. In mathematics, this state of a model is also called fixed point.
If the system variable is the average phosphorus concentration in a lake, the
concentration stays constant when all input and output processes exactly
balance each other. Since the size of some of these processes, e.g. the output,
50 LINEAR MODELS WITH ONE VARIABLE
J J
V∞ = − = , k = 0 and k = 0 (4.16)
k k
J 500 kg year−1
M∞ = − = − = 2,500 kg
k −0.2 year−1
Note: Of course, this representation of the fish pond is very simplistic.
In nature, growth and death rates are never entirely constant, so that
M ∞ would always fluctuate around an average value.
Each dynamic model im- To summarize: By setting the left side of Eq. (4.5) equal to zero, the differen-
plicitly contains a static tial equation becomes the ordinary algebraic equation (4.16). This equation
model. can also be seen as a static model for V∞ as a function of J and k . In that
sense, dynamic models that have steady-state solutions implicitly contain
a static model as well.
4 The symbol for infinity, which is used here as superscript, indicates that the
steady state is formally only reached after an infinite amount of time.
CONSTANT COEFFICIENTS 51
To model this situation, we first establish a mass balance for the substance
with the help of Eq. (4.2):
Change of mass Supply Removal Reaction
= − −
in the lake per time per time per time per time
Mathematically, the mass balance can be formulated as a differential equa-
tion (M : total mass in the lake [ M ], R: reaction per time [ M T−1 ]):
dM
= Jin − Jout − R (4.17)
dt
For the mass input Jin and the mass removal Jout we can write:
For the first-order reaction R of the substance in the lake we can write:
R = kr M (4.20)
where kr is the first-order reaction rate and has the dimension [ T−1 ]. Thus,
we end up with the following mass balance:
dM
= QCin − QC − kr M . (4.21)
dt
52 LINEAR MODELS WITH ONE VARIABLE
The quotient Q 1
V is called the specific flow rate kw . The inverse, kw , is
the renewal time τw of the lake water. The sum kw + kr is the total specific
1
rate ktot . The inverse, ktot , is the mean retention time of the substance in
the lake (influenced by both flow and decay).
Equation (4.22) can also be written in terms of the substance input Jin :
dC Jin
= − ktot C
dt V (4.23)
= jin − ktot C
The variable jin is the substance input per volume and time with the di-
mension [ ML−3 T−1 ]:
Jin Q
jin = = Cin = kw Cin (4.24)
V V
Comparing the dynamic equation for the flow reactor (4.23) with
Eq. (4.14) yields the following analogy: The system variable in the flow
reactor is C. The inhomogeneous term J corresponds to jin = kw Cin . The
coefficient k is given by the total specific rate ktot = kw + kr .
We obtain the equilibrium (steady state) concentration in the lake with
Eq. (4.16):
jin kw kw
C∞ = = Cin = Cin (4.25)
ktot ktot kw + kr
kw 1 Q 1
C∞ = Cin = Cin = Jin (4.27)
ktot ktot V ktot · V
CONSTANT COEFFICIENTS 53
The parameter p of the static model thus turns out to depend on the total
rate constant and the lake volume:
1
p= (4.28)
ktot · V
As noted before, p has the dimension [ TL−3 ].
In the following example we utilize the relation between model param-
eters and steady state:
We assume that the conditions for a linear flow reactor (one-box model)
are fulfilled. In that case, all three processes of the mass balance can be
described as linear functions:
dC
= kg C − kw C − ks C
dt (4.29)
= (kg − kw − ks )C
where:
kg = 0.5 d−1 specific growth rate of the algae
kw = Q/V flow rate through the pond
ks = 0.2 d−1 sedimentation rate of the algae in the pond
For the algae concentration not to decrease, the sum of all rates on the
right side of Eq. (4.29) must be positive or zero:
kg − kw − ks ≥ 0 (4.30)
Ties...
kw kw
C(t) = (C 0 − Cin )e−kt + Cin
k k (4.33)
= (C 0 − C ∞ )e−kt + C ∞
and assume that k is positive. The first term on the right-hand side of
Eq. (4.34) describes the exponential decay of the initial concentration C 0
in the system. For t → ∞, this term becomes zero. The old system state
C 0 is “washed out” of the system; the term describes the so-called washing-
out or decay curve. In contrast, the second term describes how—starting
from concentration zero—the system develops towards the steady state
belonging to the (constant) input jin = kw Cin . We will call this term the
ingrowth curve. Figure 4.3 shows these two components of the solution as
well as their sum. In the language of mathematics one can also say that
the general solution of Eq. (4.34) is a linear superposition of two special
solutions, the net decay curve and the net ingrowth curve.
CONSTANT COEFFICIENTS 55
C0
ln | C(t )−C ∞ |
C∞
Ingrowth curve
Decay curve
0
Time t Time t
Fig. 4.3: A linear system’s change from its initial state C 0 to the steady state C ∞
can be interpreted as the sum of two processes, the “decay” of the initial state C 0
(C 0 e−kt ) and the “ingrowth” of the end state C ∞ (C ∞ (1−e−kt )). The continuous
curve, the sum of the two dashed curves, shows the course of the system C(t).
Plotting on a logarithmic scale (on the right) gives a straight line with slope −k
Similar to Eq. (4.10) we can rewrite Eq. (4.33) in the following form:
C(t) − C ∞
ln( ) = ln(C(t) − C ∞ ) − ln(C 0 − C ∞ ) = −kt (4.35)
C0 − C∞
5 The middle part of Eq. (4.35) only makes sense if (C 0 − C ∞ ) and thus also
(C 0 − C(t)) are positive. How to modify Fig. 4.3 for the case where C 0 < C ∞ is
left as an exercise to the reader.
6 As a reminder: In the original formulation of the linear differential equa-
tion (4.5) we introduced the first-order term with a positive sign (specific rate
k ). Thus, a finite steady state exists when k < 0.
56 LINEAR MODELS WITH ONE VARIABLE
Conversely, the absolute value of the deviation between C ∞ and C(τκ ) shall
have the same value as κδC 0 :
|C ∞ − C(τκ )| = κ δC 0 = κ |C ∞ − C 0 | (4.38)
|C ∞ − C 0 | e−kτκ = κ |C ∞ − C 0 | (4.39)
C(t)
Fig. 4.4: Adjustment time
0.05 C 0 C for a linear system, drawn
0.14 C 0 for the case C 0 < C ∞ .
0.37 C 0
After elapsed time t >
τ5 % = k3 , the deviation
C0 from the steady state is
less than 5 %
C0
1/k 1/k 1/k
t
A(τ1/2 )
= 0.5 = e−kλ τ1/2
A0
ln 0.5 = −kλ τ1/2 (4.45)
ln 0.5 ln 2 0.693
τ1/2 = − = =
kλ kλ kλ
The half-life corresponds to the adjustment criterion κ = 0.5.
Until now, we calculated the adjustment time from a relative criterion (rel-
ative adjustment criterion κ): it only depends on the rate k, not on the
initial state of the system (see Eq. (4.41)). It can happen, however, that
we are interested in an absolute criterion, for instance, the condition that
the concentration should approach the steady state up to a deviation of no
more than 0.02 mg L−1 . As the following example shows, the corresponding
adjustment time then also depends on the initial concentration.
− ln 0.1 2.3
τ10 % = ≈ h = 23 h
kr 0.1
− ln 0.01 4.6
τ1 % = ≈ h = 46 h
kr 0.1
TIME-DEPENDENT COEFFICIENTS 59
In order to meet the discharge concentration limit even for the worst
case (C 0 = 10 µg L−1 ), the effluent must be pretreated in the reactor
for at least 46 h, i.e. about 2 days. Note that even in the best case, i.e.
with an effluent concentration ten times smaller, we would still need
a retention time of about 1 day or just about half the time needed for
the worst-case effluent concentration.
7 For the following discussion, we use Eq. (4.14) with the minus sign in front of
the first-order reaction, since this equation has a finite steady state for J > 0.
In principle one could also use Eq. (4.5) as the starting point.
60 LINEAR MODELS WITH ONE VARIABLE
The solution of Eq. (4.47) can be constructed with the aid of Eq. (C.8)
from Appendix C.1:
t =t
C(t) = C 0 e−kt + e−k(t−t ) jin (t )dt (4.49)
t =0
I would rather At first glance, this equation seems very complex. Yet, the first term on
have a bone than
these scrawny
the right-hand side should look familiar. In the solution of the case with
integrals! constant coefficients (Eq. 4.34) it describes the linear decay of the initial
state C 0 , that is, the decay curve (Fig. 4.3). In essence it describes the
behavior of all processes that occurred before time t = 0. In contrast,
the integral forming the second term of Eq. (4.49) constitutes the “book-
keeping” for the input from time t = 0 up to the present time t = t. This
also explains the role of the auxiliary variable t . Because we characterize
the present with the time t and look back from there, we need a second
variable t to keep track of which section of the past we are dealing with at
the moment.
In the integral, the input function jin (t) is summed up for the time
interval 0 to t. However, because the elimination process (−kC) is active
in the system as well, only a fraction of the introduced substance remains
at time t. The proportion still remaining in the system is larger the more
recent the substance input took place. The factor e−k(t−t ) describes exactly
this effect: for the current input t = t, the exponential function is 1. For
the input farthest in the past, i.e. for t = 0, the weighting is e−kt . The
influence of any input dating back even further is already contained in the
initial value C 0 .
Formally, we can therefore make the structure of the discussed integral
plausible by imagining that the continuous input is divided into many small
parts. Because the system is linear, the influence of all these small input
events can be summed up to get the overall temporal development of the
system. In mathematics this is called the linear superposition principle.
Figure 4.5 depicts this principle for two input events. The first input takes
place at time t1 with the magnitude j1 and duration Δt. Due to the input
event, the amount j1 Δt of substance per volume is added to the system.
For this first event, we can consider the system as a homogeneous system
with the initial condition C 0 = j1 Δt. The development of the concentration
for C1 (t) is then8 :
C1 (t) = j1 Δt e−k(t−t1 ) (4.50)
The second input event yields accordingly:
C2 (t) = j2 Δt e−k(t−t2 ) (4.51)
The expression (t−t2 ) in the exponential function indicates that the second
input took place at time t2 .
C1, C
C2
C1
t
t1 t2
If the continuous input is partitioned in ever more and smaller input events,
the following sum results:
n
C(t) = ji Δt e−k(t−ti ) (4.53)
i=0
Note that the input function jin (t ) can take an arbitrary form. In certain
cases, the resulting integral can be solved analytically, whereas in others
only a numerical integration is possible.
β is the rate at which the input grows each year, so for our example9
β = 0.1 year−1. The average phosphorus concentration in the lake, C(t),
can be calculated by inserting Eq. (4.55) into Eq. (4.49):
t
−kt
C(t) = C(0) e + e−k(t−t ) jin (0) eβt dt (4.56)
0
6 mg m−3 year−1
C(t = 4) = 50 mg m−3 e−4×0.12 − e−4×0.12
(0.12 + 0.10) year−1
6 mg m−3 year−1
+ e4×0.1
(0.12 + 0.10) year−1
= (30.9 − 16.9 + 40.7) mg m−3
= 54.7 mg m−3
Correspondingly, for t = 8:
Note the increasing speed at which C(t) is growing. In the first 4 years the
concentration increases by just under 5 %, while the next 4 years bring an
increase of 27 %!
e0.1 = 1.105, which is slightly larger than 1.1. The “precise” β is calculated
through the relation β = ln 1.1 = 0.0953 year−1 .
TIME-DEPENDENT COEFFICIENTS 63
Let us now take a closer look at Eq. (4.57). We observe that C(t)
depends on two specific rates. First, the development of the concentration
is driven by the already discussed total elimination rate k, which deter-
mines how quickly a substance introduced to the system is removed from
it. It is therefore a measure for the memory of the system. Second, it is
driven by β, which characterizes the speed with which the external relation
changes (in our example, the phosphorus input). Furthermore, we see that
because of the factor e−kt , the first two terms on the right-hand side of
Eq. (4.57) become smaller for increasing time and for very large times can
be neglected entirely (more precisely, they can be neglected for t k −1 ).
In contrast, the last term grows exponentially (because β > 0). For very
large times we can therefore approximate Eq. (4.57) by:
jin (0) βt jin (t)
C(t) ≈ e = , if t k −1 (4.58)
k+β k+β
jin (t) is the current input (see Eq. 4.55). This result reminds us of the
steady state (Eq. 4.48), except that it contains the additional factor β in
the denominator. The relative size of k and β therefore determines the
extent to which C(t) from Eq. (4.58) and C ∞ (t) from Eq. (4.48) differ from
each other. Let us now look at the two extreme cases:
After all this math, we want to go back to Example 4.9 and see what
happens when the input growth is stopped:
We need to examine the data from the lake more closely: The phospho-
rus load in the lake grew approximately exponentially with a rate of β =
0.1 year−1 for 20 years. The total specific elimination rate of phosphorus
in the lake is k = 0.12 year−1 . Obviously, the external perturbation of the
lake corresponds to a non-adiabatic situation.
Twenty years is long enough to apply the approximative solution
Eq. (4.58):
jin (t = 20 year)
C(t = 20 year) ≈
k+β
Because the two rates β and k have the same order of magnitude, after
20 years the lake is not in the steady state corresponding to its current
input. The hypothetical steady-state concentration for the lake would be:
jin (t = 20 year)
C ∞ (t = 20 year) =
k
The phosphorus concentration in the lake is smaller than the hypothetical
steady state concentration by:
200
is the steady state concen-
tration corresponding to
the current input. Despite
the stabilized input, the
average phosphorus con-
100 centration keeps rising for
years
Input is stabilized
0
0 10 20 30 40 50
Time [yr]
We use Eq. (4.62) and insert β = 0.02 year−1 . Because β is much smaller
than k, the third term of the equation is speed-determining. Figure 4.7
shows the calculated development of the average phosphorus concentration
C(t) and the influence of the three terms of Eq. (4.62). Terms 1 and 2
practically neutralize themselves and both tend rapidly towards zero, thus
C(t) is determined by the third term.
Term 1
C(t) [mg m
-200
Term 2
-400
0 10 20 30 40 50
Time t [yr]
As an example, we look at the same lake again. This time, however, the
remediation measures are implemented rapidly:
Figure 4.8 shows the development calculated from Eq. (4.62) and the in-
fluence of the individual terms. As expected, term 3 rapidly tends to zero No
point in
and can be neglected. Term 2 declines more slowly, but because of the fac- hurrying...
tor jin (0)/β it is much smaller than term 1 and can also be neglected.
Thus, the average phosphorus concentration in the lake is determined by
term 1. In contrast to Example 4.11, the remediation’s success is not de-
termined by the speed of input reduction but by the speed of the lake’s
response to the decreasing contamination. One could argue that an input
reduction rate β much larger than the system’s innate elimination rate
k does not accelerate the recovery of the lake and is thus economically
inefficient.
200
C(t) [mg m
100
Term 1
Term 2
0
Term 3
-100
0 10 0 30 40 50
Time t [yr]
68 LINEAR MODELS WITH ONE VARIABLE
For the sake of completeness, we will briefly look at the special case
where k = β ≡ −β. In this case, the integration of Eq. (4.56) does not lead
to Eq. (4.57), but is (with β = −k):
t
−kt
C(t) = C(0) e + e−k(t−t ) jin (0) e−kt dt
0
t
(4.65)
In this case, the characteristic rates of the system and the external relation
are identical.
Oh, periodic
disturbance!
Time
Again, let’s look at the linear flow reactor as an example. The average
concentration in the reactor under the influence of a periodically fluctuating
input is described by the following differential equation:
dC
= jin (t) − kC = j0 + j1 sin ωt − kC (4.68)
dt
The solution is calculated with Eq. (4.49) by inserting Eq. (4.67) for jin (t)
and splitting the integral into two parts:
t t
C(t) = C 0 e−kt + j0 e−k(t−t ) dt + j1 sin ωt e−k(t−t ) dt (4.69)
0 0
Let’s compare this expression with the variable steady state C ∞ (t) corre-
sponding to the momentary input j(t):
j(t) j0 j1
C ∞ (t) = = + sin ωt (4.74)
k k k
By comparing these two equations we notice the following:
k = 0.01 d−1
Figure 4.10 shows the substance input jin (upper diagram) and the
average substance concentration C(t) (lower diagram, black curve).
The grey curve describes the adiabatic approximation, that is, the
concentration that would be in equilibrium with the current input.
Because k and ω are of similar magnitudes, the adiabatic approxima-
tion does not accurately represent the real concentration in the lake.
In fact, the amplitude is smaller by a factor of 0.5 and delayed by
η = arctan( 0.0172
0.01 ) = 1.04. To convert the delay into a time, we have
to compare η with the full period of the sine function (2π = 6.28) and
multiply it with the period of variation T = 365 d:
η 1.04
ΔT = T = 365 d = 60 d
2π 6.28
TIME-DEPENDENT COEFFICIENTS 71
0.012
0.008
0.006
0.004
0.002
jin (t )
0
T
1
Concentration C (t ) [mmol m−3]
0.8
0.6
0.4
0.2
C (t )
0
0 100 200 300 400 500 600 700
Time t [d]
j0 j1 sin (ωt)
C(t) ≈ + = C ∞ (t), for kt 1 and ω k (4.75)
k k
In this case, the average concentration in the reactor follows the steady
state that corresponds to the current input j(t) = j0 + j1 sin ωt. The slow
variation of the external relation thus represents an adiabatic perturba-
tion.
1:
j0 j1 π
C(t) ≈ + sin (ωt − ), for kt 1 and ω k (4.76)
k ω 2
A linear system filters ex- The current system state C(t) lags behind the current perturbation jin
ternal fluctuations that are by a quarter period ( π2 ). The amplitude of the system fluctuation ( jω1 ) is
considerably faster than the reduced by the factor ωk 1 relative to the amplitude of the steady state
system’s reaction time. j1
( k , see Eq. 4.74). The system thus remains close to the average steady
state jk0 . In effect, a linear system filters external fluctuations out of the
system when they are substantially faster then the system’s reaction time
k −1 .
Let us demonstrate these extreme situations with the temperature
regime in a small lake.
TIME-DEPENDENT COEFFICIENTS 73
2π
Θ(t) Θeq + Θamp sin t = Θeq (t) (4.79)
T
kΘ 2π π
Θ(t) Θeq + ( )Θamp sin ( t − ) (4.80)
ω T 2
a Note that Θ
eq is not equal to air temperature since the direct exchange
of sensible heat between water and air is just one (minor) process of heat
exchange.
74 LINEAR MODELS WITH ONE VARIABLE
Since kωΘ = 0.1/6.28 = 0.016, the second term of Eq. (4.80) is fairly
small whatever the temperature fluctuation Θamp may be. Thus, water
temperature is virtually unaffected by the diurnal variability of the
relevant meteorological parameters (including the extreme day/night-
variation of short-wave solar radiation). The exact solution is plotted
in the lower panel of Fig. 4.11.
(t ) [ C]
and
eq (t )
Time t [d]
(t) [ C]
and
eq (t )
Time t [d]
Fig. 4.11: Average water temperature in a lake with annual variation (above)
and diurnal variation (below). kΘ = 0.1d−1 , Θeq = 10 ◦ C, Θamp = 10 ◦ C. The
black plot is the actual temperature Θ(t); the grey plot is the theoretical equilibrium
temperature Θeq (t). In the annual case, the actual temperature slightly lags behind
the theoretical equilibrium temperature, while in the diurnal case the fluctuation
is so rapid that the actual temperature cannot keep up and hardly deviates from
the average
QUESTIONS AND PROBLEMS 75
Question 4.3: Write down the most general form of a linear, first-order
inhomogeneous differential equation.
Question 4.5: Which dimensions do the variables in Eqs. (4.17) and (4.22)
have?
Question 4.6: What is the adjustment time of a linear model? Why is this
time not unambiguously defined, but needs the specification of a conven-
tion? Name such a convention.
an exponential function (growth rate β = 6.25 % per year). The birth rate
(b) in the city is 2.3 % and the death rate (d) is 0.5 % of the population
per year. Effects that are driven by the population’s age structure are not
taken into account. A census showed that in 1995, 10.9 million people lived
in the city, and 0.25 million migrated into it in the same year. How big
is the city’s population in the year 2020, if growth continues at the same
rate?
(a) Modify Eq. (4.77) to account for the effect of the through-flow of water
(inlets, outlet). Assume that the temperature of the inflowing water
Θin is at equilibrium with the momentary meteorological conditions
(Θin (t) = Θeq (t)) while the temperature of the outflow is equal to
the mean lake temperature (Θout (t) = Θ(t)). The water through-flow
is expressed by the specific flow rate kw (Eq. 4.22).
(c) Compare the result obtained in (b) with Lake B, which has the same
thermal properties but no through-flow. Calculate ηB and ΔTB .
(d) Assume that as in Eqs. (4.78) and (4.79), the annual temperature
variation in the two lakes can be described by a sinusoidal curve.
Calculate the ratio of the temperature amplitudes of the two lakes,
Θamp,A /Θamp,B . Explain the result qualitatively.
(a) Draw a mass balance by using a box diagram and establish the dy-
namic equation for the dye concentration in the fountain.
(c) Assume that the fountain is operated in a closed loop, i.e., the in-
flowing water is entirely made up of outflowing water. What is the
corresponding dynamic equation? Again, calculate the time tcrit for
this flow regime until the water appears uncolored.
(a) Draw a box diagram for the mass balance and establish a dynamic
equation for the concentration of substance A in the sewage plant.
(b) Give a mathematical relation for the degradation rate R [mg L−1
d−1 ] of substance A as a function of the concentration in the plant C.
(d) Starting from the steady state Cin = 80 mg/L and Cout = 4.0 mg/L,
how long does it take for the concentration C in the plant to fall to
5 % of its initial value, if at time t = 0, Cin suddenly becomes zero?
(e) What is the answer to question (d) if you start from the steady state
with Cin = 20 mg/L and Caus = 1.0 mg/L? Explain your answer.
(a) Establish the differential equation for the change in copper concen-
tration C in the soil.
(b) How large is the copper concentration in the soil at steady state?
(c) How long does it take until the concentration comes to within 5 % of
the steady state calculated in (b), if we assume an initial concentra-
tion of 0 mg copper per square meter?
QUESTIONS AND PROBLEMS 79
So far we have only dealt with models described by a single system variable.
We called these models one-dimensional. To adequately describe a system,
however, we often need several variables that interact with each other. This
We dogs demand
is where multi-dimensional models come into play. NOT
There are two ways to move from a one-dimensional to a multi-dimen- to be thrown
sional model. First, it could be useful to divide the system variable V into into the same
box as
subcomponents. For instance, we want to split the total phosphorus con- humans!
centration in a lake into dissolved and particle-bound phosphorus. Or we
are interested in two substances A and B that can be converted into each
other by a chemical transformation. In both cases we need to add material
differentiation to our model.
Second, spatial differentiation could be necessary in some scenarios, for
instance, dividing the overall average phosphorus concentration in a lake
into surface and deep water concentration. In this case, we call our model
a two-box model, because it contains two spatially separate “boxes”. Of
course, this expression could also be used for the materially differentiated
model: the boxes would be material compartments rather than spatial ones.
Mathematically speaking it makes no difference whether we differentiate
spatially or materially. Both cases lead us to a system of two or more
coupled differential equations.
As we will see, the coefficient matrix P plays a central role in the solution of
a system of linear differential equations (even more so in systems with more
than 2 dimensions). Therefore, we want to introduce four characteristic
quantities of the matrix P (see Appendix D):
1. The determinant of P:
det(P) = p1,1 p2,2 − p1,2 p2,1 (5.7)
2. The trace of P:
tr(P) = p1,1 + p2,2 (5.8)
84 LINEAR MULTI-DIMENSIONAL MODELS
3. The discriminant of P:
1
λi = tr(P) ± Δ(P) , i = 1, 2 (5.10)
2
4 “Eigen” is a German word that means “own”, used in the sense of the char-
acteristic (own) values belonging to a particular matrix.
TWO-DIMENSIONAL MODELS 85
The four other constants ai,j in Eq. (5.12) are better legible if we don’t
express them by R1 and R2 , but rather by the constants a1,0 and a2,0 of
Eq. (5.13) (for λ1 , λ2 real and negative and λ1 = λ2 ):
a1,1 = q· (p1,1 − λ2 )(V01 − a1,0 ) + p1,2 (V02 − a2,0 )
a1,2 = −q· (p1,1 − λ1 )(V01 − a1,0 ) + p1,2 (V02 − a2,0 )
a2,1 = q· p2,1 (V01 − a1,0 ) + (p2,2 − λ2 )(V02 − a2,0 ) (5.14)
a2,2 = −q· p2,1 (V1 − a1,0 ) + (p2,2 − λ1 )(V2 − a2,0 )
0 0
− 1
q = 1
λ1 −λ2 = √ 1
= (p1,1 − p2,2 )2 + 4 p1,2 p2,1 2
Δ(P)
Only in rare cases would we calculate the solution of Eq. (5.1) manually.
It makes more sense to use computer software for that. However, we can
get a qualitative view of the system’s behavior from the eigenvalues of the
coefficient matrix P. They are, as Eq. (5.10) shows, relatively easy to find,
at least for a two-dimensional system.
If the reactor is run with constant conditions for long enough, the concen-
trations CA and CB in the reactor (and therefore also in the outflow) take
∞ ∞
on the steady state values CA and CB . The conversion constraint of 90 %
can be written as:
∞
CB 1
0.9 ≤ ∞ ∞ = ∞ ∞
CA + CB CA /CB + 1
∞ ∞
Solved for CA /CB :
∞ ∞
CA /CB ≤ (1/9) = 0.11
86 LINEAR MULTI-DIMENSIONAL MODELS
∞ ∞
Before we use Eq. (5.13) to calculate CA and CB , we have to ensure
that the eigenvalues are real and negative and that the determinant is
det (P) = 0. According to Example 5.1, we have:
det (P) = (kA + kw )(kB + kw ) − kA kB
2
= kw + kw (kA + kB ) > 0
This is because all coefficients kw , kA , kB are larger than zero. Furthermore,
you can verify that all eigenvalues correspond to case 1 in Table 5.1. Thus
∞ ∞
we can now calculate the proportion CA /CB from Eq. (5.13), making use
of the fact that R2 = 0 (see Example 5.1):
∞
CA −p2,2 R1 kB + kw
∞ = = ≤ 0.11
CB p2,1 R1 kA
Solving for kw = Q/V gives:
Q
kw = ≤ 0.11kA − kB
V
= 0.11 × 0.5 h−1 − 0.01 h−1
= 0.045 h−1
or
Q ≤ 10 m3 × 0.045 h−1 = 0.45 m3 h−1
So the flow rate Q must not increase beyond 0.45 m3 h−1 .
tr < 0, Δ > 0
(1) Real, both negative Solution with steady state
0 < Δ1/2 < −tr
tr < 0, Δ > 0
(2) Real, λ1 = 0, λ2 < 0 Δ1/2 = −tr Solution with steady state
(corresponds to det = 0)
Solution → ∞
(3) Real, at least one positive tr > 0, Δ > 0
(solution unbounded)b
Conjugate-complex,
(5) tr < 0, Δ < 0 Damped oscillation
with negative real part
Conjugate-complex,
(6) tr > 0, Δ < 0 Oscillation → ±∞
with positive real part
a
To simplify notation, we don’t write the argument (P) in Δ, tr und det. In addition, the case Δ = 0(λ1 = λ2 ) is not
TWO-DIMENSIONAL MODELS
For the model in Fig. 5.3, the model parameters pi,j from Eq. (5.1) take the
following values:
One can confirm that under these conditions, the coefficients in matrix P
always have the following properties:
(a) The elements outside of the diagonal p1,2 and p2,1 are positive. The
discriminant Δ of the characteristic equation is therefore larger than
zero (Eq. 5.9) and the eigenvalues are real.
(b) The diagonal elements p1,1 and p2,2 are negative. Furthermore,
−p1,1 ≥ p2,1 and −p2,2 ≥ p1,2 . The eigenvalues are therefore neg-
ative (and real). The model tends to the steady state. Its adjustment
time τ is determined by the smaller of the eigenvalues. Analogously
to the one-dimensional system, the 5 % adjustment time is:
3 6
τ5 % ≈ = (5.16)
min |λi | −tr(P) − Δ(P)
for λi = 0, Δ(P) > 0 and tr(P) < 0. However, Eq. (5.16) cannot be
interpreted with the same stringency as its one-dimensional counter-
part (Eq. 4.42). Individual variables of a multi-dimensional system of
equations can approach their steady state much more quickly if the
coefficients ai,j belonging to the smallest eigenvalue are small or even
zero for the respective variable. Therefore, Eq. (5.16) should always be
used merely as a first, rough estimate. A more detailed examination
would require knowing the explicit solution.
(c) If the system is conservative with respect to its mass balance, we
have a special case: there are neither internal sources nor sinks. This
90 LINEAR MULTI-DIMENSIONAL MODELS
means that the sum of both variables is constant and one of the
two eigenvalues is zero. For the coefficient matrix, −p1,1 = p2,1 and
−p2,2 = p1,2 . The first row of matrix P is identical with the second
row except for the algebraic sign. Such a matrix is called singular.
Its determinant is zero: det(P) = 0. Singular matrices have at least
one eigenvalue that is zero.
det(P) = (0.54 × 0.05) h−2 − (0.01 × 0.5) h−2 = 2.2 × 10−2 h−2
tr(P) = −(0.54 + 0.05) h−1 = −0.59 h−1
Δ(P) = tr(P)2 − 4 det(P) = 0.260 h−2
1
λi = [−0.59 h−1 ± (0.260)1/2 h−1 ]
2
λ1 = −0.040 h−1 ; λ2 = −0.55 h−1
From any given initial state, the overall 5 % adjustment time according to
Eq. (5.16) is:
3 3
τ5 % = = = 75 h
|λ1 | 0.04 h−1
This time is about equal to the average retention time of the water in
the reactor (τw = V /Q = 25 h; τ5 % ∼ 3τw ). It depends only marginally
on the time for reaching an equilibrium between phases A and B. As
shown in Example 5.4, the latter depends on the sum of the two rates
(kA + kB )−1 ≈ 2 h.
In the next example, we look at a system with a singular system ma-
trix, i.e. det(P) = 0. Such systems only have a steady state if they are
homogeneous, that is, if R1 = R2 = 0.
TWO-DIMENSIONAL MODELS 91
AH + H2 O H3 O+ + A−
or shorter:
k1
M1 M2
k2
Instead of one specific steady state or fixed point, we get a whole set of
potential steady states which all lie on a straight line, the so-called fixed-
point line. It corresponds to the thermodynamic equilibrium of the reaction
with the equilibrium constant K = kk12 .
We don’t need complex calculations to determine which point on the
fixed-point line the model is actually moving towards. We only need to
remember that the model obeys a mass conservation condition. Figure 5.4
shows the phase diagram of the homogeneous equilibrium reaction (5.17).
On the one hand, mass conservation means that the system can only move
0
along the line with slope (−1), which crosses the M1 or M2 axis at Mtot =
0 0
M1 + M2 . This is the system’s trajectory, labeled A in Fig. 5.4. On the
other hand, Eq. (5.21) says that all steady states must lie on a line through
the origin with slope ( kk12 ). This is the system’s fixed-point line, labeled B in
Fig. 5.4. The steady state in accordance to the initial condition (M10 , M20 )
lies at point S, the intersection between the two lines.
What remains is the confirmation that the model does indeed move towards
the fixed-point line. For this, we solve the system of equations (5.17) by
replacing M2 with Mtot0
− M1 in the first equation. We get the following
for M1 :
dM1 0
= k2 Mtot − (k1 + k2 )M1 (5.22)
dt
Eq. (5.22) is an inhomogeneous linear first-order differential equation with
the variable M1 . We already know its solution from Chap. 4 with Eqs. (4.6)
and (4.33):
M1 (t) = M1∞ + (M10 − M1∞ ) e−(k1 +k2 )t (5.23)
0
The solution for M2 (t) results either from M2 = Mtot − M1 or analo-
gously to Eq. (5.23):
dN1
= −kλ,1 N1
dt (5.27)
dN2
= kλ,1 N1 − kλ,2 N2
dt
kλ,i is the decay constant of the isotope i. It is linked to the half-life
τ1/2,i through the following relation: kλ,i = ln 2/τ1/2,i .
We will now examine this system of equations more closely. We notice that
N2 depends on N1 , but N1 doesn’t depend on N2 . Such a system is called
hierarchical. The hierarchy of the system is expressed in the so-called
94 LINEAR MULTI-DIMENSIONAL MODELS
triangular form of its coefficient matrix, in which all elements above the
diagonal are zero6 :
−kλ,1 0
P= (5.28)
kλ,1 −kλ,2
This makes determining the eigenvalues and therefore the solution of the
system of equations particularly easy. From the characteristic equation
(5.11) it follows that the eigenvalues are identical with the diagonal
elements:
If we now insert the solution for N1 into the second differential equation,
we get:
dN2
= kλ,1 N10 e−kλ,1 t − kλ,2 N2 (5.30)
dt
To solve this differential equation we can use the solution for the linear
one-box model with exponentially varying input. To do so we can view the
first term on the right-hand side of the equation as the input function of
the system variable N2 , which exponentially falls with kλ,1 . The initial value
of the input function is j0 = kλ,1 N10 . If we now apply Eq. (4.57), and with
the initial condition N20 = 0, we get:
kλ,1 kλ,1
N2 (t) = N 0 e−kλ,1 t − N 0 e−kλ,2 t
kλ,2 − kλ,1 1 kλ,2 − kλ,1 1
(5.31)
kλ,1
= N 0 · (e−kλ,1 t − e−kλ,2 t )
kλ,2 − kλ,1 1
We see that the solutions for N1 (t) and N2 (t) are sums of exponential
functions again. The exponential coefficients are the eigenvalues of the co-
efficient matrix. Because the system is hierarchical, the second exponential
function (with kλ,2 ) does not appear in the solution of N1 (t). Only the so-
lution of N2 (t) depends on both decay constants or eigenvalues. Finally, we
learn that for the homogeneous system Eq. (5.27) with negative eigenvalues,
the steady state is zero.
Let us summarize what we have learned from this example. A hierar-
chical system can be recognized by the triangular form of the coefficient
matrix. Therefore, the eigenvalues are identical with the diagonal elements
of this matrix and the system of equations can be solved iteratively. In
the next example, we will explore a hierarchical system consisting of two
spatially separate boxes:
First, we establish the mass balance for the two boxes: Mi is the total mass
of the isotope in lake i. The external isotope input is given by the input
function Ji [ MT−1 ], and the radioactive decay is described by −kλ Mi . The
transport processes we take into account are the flow from the first into the
second lake and the second lake’s outflow. We end up with the following
system of differential equations:
dM1 Input into Flow into Radioactive
= − −
dt lake 1 lake 2 decay
= J1 − Q1,2 C1 − kλ M1
96 LINEAR MULTI-DIMENSIONAL MODELS
dM2 Input into Outflow from
= + (5.32)
dt lake 2 lake 1
Outflow from Radioactive
− −
lake 2 decay
= J2 + Q1,2 C1 − Q2 C2 − kλ M2
Again, Ci = Mi /Vi means the concentration in lake i. For the concentra-
tion, the dynamic equations are:
dC1 J1 Q1,2
= − C1 − kλ C1
dt V1 V1
J1 Q1,2
= −( + kλ )C1
V1 V1
(5.33)
dC2 J2 Q1,2 Q2
= + C1 − C2 − kλ C2
dt V2 V2 V2
J2 Q1,2 Q2
= + C1 − ( + kλ )C2
V2 V2 V2
The coefficient matrix has the form:
⎛ ⎞
Q
− V1,2 + kλ 0
P=⎝ 1
Q1,2
⎠ (5.34)
V2 − Q2
V2 + kλ
with
J2 + Q1,2 C1∞
C2∞ = (5.39)
Q2 + V2 kλ
To illustrate this rather complex-looking result, we can look at the concrete
case where the outflow of a relatively large lake leads into a much smaller
lake. Maybe you can already intuitively guess how a radioactive substance
with half-life τ1/2 of about 12 year (this could be tritium, a radioactive
isotope of hydrogen denoted 3 H) behaves in this system:
We can now calculate the steady-state concentrations from Eqs. (5.37) and
(5.39):
C1∞ = 33 Bq m−3
C2∞ = 60 Bq m−3
Let’s assume that the isotope influx (J1 and J2 ) is stopped abruptly. Of
course, this assumption is unrealistic, because the lakes’ catchment area
also has a “memory”, for instance in the soil, where the isotope is washed
out only slowly. Nevertheless, we can hypothetically examine how C1 and
C2 (the concentration in the two lakes) would develop.
First, we can see that the only change in the system of equations (5.33)
concerns the two inhomogeneous terms (J1 = J2 = 0). The system matrix
P and therefore also the eigenvalues don’t change. They have the following
values:
λ1 = −0.09 year−1 ; λ2 = −0.96 year−1
As per Eq. (5.16), the (absolute) smaller eigenvalue λ1 determines the lake
chain’s reaction to the input change:
3
τ5 % = = 33 year
0.09 year−1
In fact, the two eigenvalues are each composed of the water flow rate and the
decay constant. In this example they can therefore explicitly be allocated
to the two subsystems. The eigenvalue λ1 describes the isotope in the first
lake, λ2 the one in the second lake. Figure 5.6 shows the reduction of
98 LINEAR MULTI-DIMENSIONAL MODELS
40
30
Lake 1
20
10 Lake 2
0
0 5 10 15 20 25 30 35 40
Time t [yr]
tritium in the two lakes. Starting from the steady-state concentration, the
tritium concentration in the second lake declines very rapidly. The first lake,
however, reacts more slowly. If the second lake did not lie downstream of
another lake, its adjustment time would be τ5 %,Lake2 = 3/0.96 ≈ 3 year.
The memory of the downstream lake is thus located in the water of the
upstream lake, much in the same way as the memory of both these lakes
lies in the soil of their watersheds. In Problem 5.2 we will return to this
example.
dC
= kw Cin − kw Cout − kr C (5.40)
dt
Cout [ ML−3 ] Average concentration in the outflow
The price for this extension is the appearance of a new system variable
Cout . This begs the question how it is related to C. As a first attempt,
we set:
Cout
= γ = const. (5.41)
C
γ [-] Stratification parameter for phosphorus
By doing so we can now eliminate the system variable Cout in Eq. (5.40)
and again get a one-box model. This model is different from Eq. (4.22):
dC
= kw Cin − (γkw + kr )C (5.42)
dt
Epilimnion
Thermocline
Hypolimnion
Lake depth
Water Water
temperature density
Fig. 5.7: Vertical temperature and density profile of a lake during summer. The
thermocline separates the warm surface water, called epilimnion, from the cold
deep water, called hypolimnion. In the thermocline, vertical water and mass trans-
port is reduced in comparison to the layers above and below
time t, that is, as C(z, t). The resulting equations, however, are usually so
complex that they can only be solved numerically by computer.
Multi-box models stand It is not always justified to jump straight from using a one-box model
as an intermediate step to a continuous space-time-model. Often, an intermediate solution is most
between one-box models practical. Figure 5.7 suggests such a solution: the partition of the lake
and spatially continuous into two subsystems, the epilimnion and the hypolimnion, that can each
models. individually be treated as completely mixed. Similarly, in describing the
atmosphere, we could divide it into layers such as the troposphere and
stratosphere. This approach can deliver acceptable results for many types
of systems.
As Fig. 5.8 shows, the spatial segmentation of a system entails the ques-
tion of how to describe the mass flow between the different subsystems. We
will discuss this problem more extensively in Sect. 8.2. For now, and in the
following example, we will anticipate and use the outcome of that discussion
without further explanation.
First, we establish the mass balance for the two boxes. For the epilimnion,
it is:
dME
= {Inflow} − {Outflow} − {Decay} ± {Exchange}
dt (5.45)
= QCin − QCE − kr,E ME − Qex CE + Qex CH
After dividing Eq. (5.45) by the epilimnion volume VE and Eq. (5.46) by
the hypolimnion volume VH we get the dynamic equations for the concen-
trations:
dCE Q Q Qex Qex
= Cin − CE − kr,E CE − CE + CH
dt VE VE VE VE
= kw Cin − kw CE − kr,E CE − kex,E CE + kex,E CH
= kw Cin − (kw + kr,E + kex,E )CE + kex,E CH
(5.47)
dCH Qex Qex
= CE − CH − kr,H CH
dt VH VH
= kex,H CE − kex,H CH − kr,H CH
= kex,H CE − (kex,H + kr,H )CH
102 LINEAR MULTI-DIMENSIONAL MODELS
Table 5.2: Numeric example for the mass balance of a stratified lake
Steady states:
∞
CE = 73 mg m−3 and ∞
CH = 52 mg m−3
1
λ1 = (tr(P) + Δ(P))
2
1 √
= (−0.0438 + 0.00109) d−1 ≈ −0.0054 d−1
2
1
λ2 = (tr(P) − Δ(P))
2
1 √
= (−0.0438 − 0.00109) d−1 ≈ −0.038 d−1
2
The absolute smallest The adjustment time can be estimated with the (absolute) smallest eigen-
eigenvalue generally de- value (Eq. 5.16):
termines the adjustment
time of a linear system. 3 3
τ5 % ≈ = = 555 d
|λ1 | 0.0054 d−1
The lake would therefore have to remain stratified for more than a year for
the calculated equilibrium concentrations to be reached. As we have seen,
this is not what happens in reality.
Let us try to capture the complex mixing cycle of a lake in a temperate
European climate with a model. To do so, we assume that the lake is
completely mixed from the beginning of January until the end of April,
that is, for 120 days. Then, a stagnation period of 8 months begins, during
which the lake is stratified. The model of the lake thus changes back and
forth from a one-box model in winter to a two-box model in summer. The
initial concentrations at the beginning of the two-box phase are identical
0 0
(CE = CH ), and correspond to the final concentration of the preceding
one-box phase. At the transition from the two-box to the one-box phase,
the initial concentration is calculated from the volume-weighted final state
of the two-box phase:
CE (t) VE + CH (t) VH
C0 = (5.52)
VE + VH
Figure 5.9 shows the temporal variation of the dissolved substance de-
scribed in Example 5.8 in a lake undergoing an annual cycle of stratifica-
tion and destratification. For the stratification period, the numbers from
TWO-DIMENSIONAL MODELS 105
Table 5.2 were used. For the mixing period, the elimination rate kr and the
flow rate kw were adjusted as follows:
kr,E VE + kr,H VH
kr = (5.53)
VE + VH
Q
kw = (5.54)
VE + VH
The computation starts at time t = 0 with the onset of stratification. Con-
centrations in the epilimnion and the hypolimnion develop differently. In
the epilimnion, the concentration increases, while it decreases in the hy-
polimnion, yet the steady state is not reached. After 8 months (245 days),
the lake is instantaneously and completely mixed, resulting in an average
concentration in the entire lake which lies below the steady-state concen-
tration of the mixing period. Thus, the concentration increases further over
the next 120 days but is still well below the steady-state value when the
cycle starts over at the onset of stratification.
In reality, the transition between complete mixing in winter and com-
plete stratification in summer is not sudden, but gradual. Nevertheless,
the basic outcome of the mixing cycle is reproduced quite well with this
two-phase model.
7 The relation between the descriptions in Eqs. (5.59) and (5.60) can be cal-
culated from Euler’s formula, eiωt = cos ωt + i sin ωt. If the variables yi (t) are
real functions, this also holds for the coefficients bi,j . The ai,j from Eq. (5.59)
however are generally complex numbers. For details see Appendix C.4.
TWO-DIMENSIONAL MODELS 107
To determine the two other constants bi,j , we insert Eq. (5.60) into the
differential equations (5.55) and get the identities:
dy1
= ωb1,2 cos ωt = k1 (cos ωt + b2,2 sin ωt)
dt
dy2
= −ω sin ωt + ωb2,2 cos ωt = −k2 b1,2 sin ωt
dt
which can only be fulfilled if b2,2 = 0 and b1,2 = ω/k2 = (k1 /k2 )1/2 .
We thus get this solution for our chosen initial condition:
k1
y1 (t) = sin ωt
k2 (5.62)
y2 (t) = cos ωt
In Fig. 5.10, the two functions are depicted as a time and a phase diagram
for the case kk12 = 4. The time diagram shows two undamped harmonic
oscillations that are shifted by a fourth of a period. The phase diagram
shows a closed elliptic trajectory. The system circulates the steady state
without ever reaching it.
Undamped oscillations
Undamped
collision!
Undamped
oscillations.
Note that mathematicians like to use the parameter ε whenever they want
to point out that this term can be arbitrarily small. As our analysis will
show, it is the mere existence of this additional term—and not its size—
which is responsible for a fundamental difference between Examples 5.9 and
5.10. The model has the coefficient matrix:
0 k1
P= (5.64)
−k2 −ε
Its characteristic equation is slightly modified compared to Eq. (5.57):
λ2 + λε + k1 k2 = 0 (5.65)
with the eigenvalues
1 ε ε2
λi = −ε ± ε2 − 4k1 k2 = − ± − k1 k2 (5.66)
2 2 4
TWO-DIMENSIONAL MODELS 109
Structural instability
Conducting
test sequence
for structural
stability...
110 LINEAR MULTI-DIMENSIONAL MODELS
b
1.0 1.5
1.0
0.5
0.5
0 0
-0.5
-0.5
-1.0
-1.0 -1.5
0 0.5 1.0 1.5 2.0 2.5 3.0 -1.5 -1.0 -0.5 0 0.5 1.0 1.5
If ε increases, but the condition ε < 4k1 k2 still holds, the expression under
the root in Eq. (5.67) remains negative, i.e. the eigenvalues remain non-
real. However, the angular frequency ω (which, for weak damping, remains
practically the same as in the undamped case) now drops to lower values:
ε2
ω→ω =
k1 k2 − (5.69)
4
A radical change of behavior takes place for
ε2
≥ k1 k2 (5.70)
4
According to Eq. (5.66), both eigenvalues are now real and negative and the
system “creeps” without oscillation towards the steady state y1 = y2 = 0
(see Fig. 5.11b). As we will see in Example 5.11, the equations introduced in
Examples 5.9 and 5.10 can be used to describe an oscillator or a pendulum,
among other things. Therefore, a system with strong damping is also called
a creeping oscillation or a creeping pendulum.
As a concrete example for a model with a structural instability, we use
a classic example from physics: the harmonic oscillator. In Chap. 6, we will
encounter another application from an entirely different field.
TWO-DIMENSIONAL MODELS 111
K = −f x (5.71)
d2 x
M = K = −f x (5.72)
dt2
x [L] Displacement from equilibrium position
f [ MT−2 ] Spring constant (force per displacement)
M [M] Mass
K [ MLT−2 ] Spring force
As we might remember from school, the solutions of Eq. (5.72) are har-
monic oscillations. To demonstrate the connection to the previous model
(Eq. 5.55), we choose the following new variables to describe the harmonic
oscillator:
y1 ≡ x
dx (5.73)
y2 ≡
dt
If we differentiate y1 with respect to time t, it follows that:
dy1 dx
= = y2 (5.74)
dt dt
On the other hand, if we convert Eq. (5.72) with the relation
d2 x dy2
= (5.75)
dt2 dt
then it follows that:
dy2
M = −f y1 (5.76)
dt
or:
dy2 f
= −ω y1
2
with ω= (5.77)
dt M
The dynamic equation (5.72) for the harmonic oscillator is a second-order
differential equation. With the conversions above, we have transformed it
into a system of two first-order differential equations: (5.74) and (5.77).
This system has the following coefficient matrix:
0 1
P= (5.78)
−ω 2 0
112 LINEAR MULTI-DIMENSIONAL MODELS
This matrix has the same type of eigenvalues as Eq. (5.55). Therefore, the
Watch out!
Harmonic model of the harmonic oscillator must be structurally unstable as well. If
entities are
structurally
we introduce a very small damping term (γ: damping constant) that is
unstable! proportional to the speed of the moving mass M , dx
dt , we get:
d2 x dx
M = −f x − γ (5.79)
dt2 dt
dy1
= y2
dt (5.80)
dy2 γ
= −ω 2 y1 − y2 = −ω 2 y1 − εy2
dt M
The damping constant per mass, γ/M , takes over the role of ε in Eq. (5.63).
The amplitude of the oscillation becomes smaller and smaller with time
t, until the oscillator eventually remains still at its equilibrium position
(x = 0). If the damping exceeds a certain value, the oscillator “creeps” into
this position. We will return to this type of differential equations when we
look at two-dimensional nonlinear systems in Chap. 6.
Let us now summarize our insight so far. To examine the behavior of a
model and its steady state, we need to calculate the eigenvalues λi of the
coefficient matrix. If both of them are real and negative or zero, the model
reaches a steady state. If non-real eigenvalues occur, the model still reaches
a steady state as long as their real parts are negative or zero. If purely
imaginary eigenvalues occur, the model undergoes an undamped oscillation.
Eigenvalues with positive real parts lead to solutions that diverge towards
infinity.
Multi-box model
A multi-box model at
its steady state: What
Everything flows... goes in, comes out.
It s that simple.
MULTI-DIMENSIONAL MODELS 113
of equations that are then solved with methods called linear programming.
114 LINEAR MULTI-DIMENSIONAL MODELS
300
250
0 500 1000 1500 2000
Year
Table 5.3: Exchange rates and dynamic equations of the global linear carbon
model. Labeling of fluxes and rates: first index = number of the destination box,
second index = number of the originating box
We calculate the specific exchange rates ki,j between land and atmo-
sphere as follows:
F1,2 110 × 1015 g a−1
k1,2 = = = 0.061 a−1
M2 1, 800 × 1015 g
F2,1 110 × 1015 g a−1
k2,1 = = = 0.183 a−1
M1 600 × 1015 g
Table 5.3 shows the calculations for the transfer rates and the transport
equations of the entire model. The smallest rates are, as we already noted,
the ones between the ocean surface and the deep sea.
MULTI-DIMENSIONAL MODELS 117
How can this simple model be used to answer the three questions
posed above? The dynamic development of the four carbon reservoirs Mi (t)
can be calculated analytically or with the aid of a computer program.
In the first case, each variable Mi (t) is described by a constant and four
time-dependent exponential functions (see Eq. 5.83). The initial values of
Mi (i = 2, 3, 4) are those in Fig. 5.12, while the atmospheric reservoir (M1 )
is increased by the anthropogenic input of 300 × 1015 to 900 × 1015 g C.
Figure 5.15 shows the result of a numerical simulation of the model’s
development through time. As expected, the biggest reservoir, M4 (the
deep sea), reacts most slowly. That is because it’s only coupled to the
other reservoirs by a fairly small (relative to the reservoir size) mass flux.
The second question, regarding the new steady state, can easily be an-
swered from Table 5.3 without solving the system of equations. Because we
are dealing with a linear homogeneous model, the relative sizes of the Mi at
118 LINEAR MULTI-DIMENSIONAL MODELS
λ1 = −0.443 year−1
λ2 = −0.114 year−1
λ3 = −0.00971 year−1
λ4 = 0
3 3
τ5 % ≈ = ≈ 310 year (5.85)
min (|λi | = 0) 0.00971 year−1
Of course, this carbon model is not very realistic. First, we treated the
anthropogenic disturbance as a single input event. Figure 5.13 however
shows that an exponentially growing input function comes much closer to
reality. A second simplification is the assumption that all fluxes are linear
functions of the reservoir sizes.
Nevertheless, the model illustrates important properties of the carbon
cycle that are confirmed by more complex models. As mentioned above,
the anthropogenic disturbance, although relatively small compared to the
natural fluxes, has a large impact on the atmosphere for two reasons: the
10 Of course, this leaves aside the increasing acidity of our oceans as a result of
their atmospheric carbon uptake, a process which has severe effects of its own.
MULTI-DIMENSIONAL MODELS 119
Question 5.9: For a car, the parts that behave like a creeping pendu-
lum play an important role. If they lose this property, they urgently need
replacement. Which parts are they?
Question 5.10: How can the overall reaction rate of a linear model be
determined? Are there special cases where some system variables move
towards a steady state much more rapidly than the overall rate would
indicate? Give a mathematical explanation and a concrete example.
Question 5.12: The net flux between two neighboring spatial boxes can
be described as a linear function of the concentration difference between
them. In Example 5.8 we applied this principle. What idea is it based on?
QUESTIONS AND PROBLEMS 121
The dots indicate that we are neglecting certain intermediate products with
a very short half-life. Although 210 Pb is not stable either, it has a much
larger half-life than the other isotopes shown in the chain. Its decay plays
no role in the following problem.
To simplify, we choose the following notation (half-life in parentheses):
X≡ 222
Rn (3.8 d), Y ≡ 214
Pb (26.8 min), Z≡ 214
Bi (19.8 min)
Establish the system of differential equations for the activities Ai and cal-
culate the coefficient matrix. Then, calculate the activity of 214 Bi as a
function of time, AZ (t), and from that, derive a an approximate formula
for t > 10 h. The initial activities A0i are: A0X = 1, 000 Bq, A0Y , A0Z = 0 Bq.
(1 Bq = 1 decay per second).
QF QF
Digester
V2 A2
(a) Establish the dynamic equations for the tritium activities in the com-
bined sedimentation/aeration tank (A1 (t)) and in the anaerobic di-
gestion tank (A2 (t)) for the time after the accident (t > 0).
(b) The half-life of tritium is t1/2 = 12 year. Compare the decay constant
of tritium with the different water transport rates of the system. How
large is the influence of radioactive decay on the tritium activity in
the plant’s outflow?
(c) Estimate the eigenvalues of the system. Hint: set the absolutely small-
est matrix element and the radioactive decay constant of tritium equal
to zero.
(d) With what rate does the activity in the outflow decline after a few
months?
(c) How do the answers to question (b) change if the substance is ra-
dioactive with a decay constant of kλ = 1 × 10−2 d−1 ?
(d) Answer question (b) for a non-radioactive (that is, conservative) sub-
stance which adsorbs on particles and then gets transported with
them from the surface into the deep water and finally to the sediment.
The relevant flows are linear functions of the masses M1 = C1 V1 and
M2 = C2 V2 :
Transport V1 → V2 : ks1 M1
Transport V2 → sediment : ks2 M2
Use ks1 = 0.02 d−1 , ks2 = 0.005 d−1.
Also, neglect the fact that the lake’s cross section decreases with
depth, i.e., assume that the entire sediment flux from V1 ends up in
the volume V2 . Remember that at the transition from equations for
Mi to those for Ci the volumes are different: not the same amount of
mass leaves V1 per volume as arrives in V2 per volume (Eqs. 5.45–5.48
may be helpful here).
(e) Compare the results from (b) to those from (d) and attempt to explain
qualitatively the respective ratios between C1∞ and C2∞ .
(c) After 5 days, measurements show that the concentration in Lake 2’s
outflow is decreasing exponentially (e−αt ). How large is α?
(d) How would C1 (t) and C2 (t) change if, for storing electric energy, water
would be pumped at an average rate of Qp = 4 × 105 m3 d−1 from
Lake 2 to Lake 1, and the flow from Lake 1 to Lake 2 would increase
correspondingly by Qp ? When do C1 and C2 decrease below 10 µg/L?
124 LINEAR MULTI-DIMENSIONAL MODELS
(e) Explain the differences between the results from (a) and (d)
qualitatively.
(a) Formulate the set of differential equations for x and y and discuss the
properties of the solutions. Is a steady state reached?
(b) Discuss the limits of the model. Hint: in some cases the position
of the valve may become negative. What happens qualitatively to
the solution if the dynamic equations are modified such that dx/dt
becomes 0 as soon as x drops below 0?
(c) How should the coefficients k1 and k2 be chosen to make the oscilla-
tions of y small?
(d) Are there other means to dampen the temperature fluctuations in the
building?
Chapter 6
Nonlinear Models
Nonlinear
box model, Nice, keep
right? it up!
Natural systems are rarely linear. Even if they are, that linearity often
exists only within a limited range of the system variables Vi . Indeed, the
surprising diversity in the behavior of natural systems is mostly a result
nonlinear processes. Despite ever more powerful computers, the behavior
of many nonlinear systems can only be predicted over small periods: the
weather forecast is a prime example of this.
Continuous nonlinear systems (in time) are described by nonlinear dif-
ferential equations. If the system variables are continuous in time and space,
we are dealing with nonlinear partial differential equations, such as those
occurring in fluid dynamics. The field of fluid dynamics describes the dy-
namics of gaseous and fluid systems and forms the basis of atmospheric
science and oceanography. It can be used to model such phenomena as the
Gulf stream or a tropical cyclone. In this chapter, we will start with spa-
tially discrete models, i.e. our usual box model approach. In Chap. 8 we
will introduce spatially continuous models and get a first glimpse at the
colorful world of these types of models.
Nonlinear differential equations can only be solved analytically in spe-
cial cases. We usually have to apply numerical methods, which is not the
topic of this book. In some cases, however, nonlinear models can be ap-
proximated by linear equations so that their behavior near the steady state
can be analyzed. But in this chapter we will also go beyond such approxi-
mations and discuss some behavioral traits of nonlinear models that have
no resemblance with the behavior of their linear siblings anymore.
dV
= g(V(t), t) (6.1)
dt
With this seemingly cumbersome notation we want to point out that the
general function g can depend on time both explicitly and implicitly via V
(see Sect. 4.1). The explicit time dependence arises from the external and
the implicit time dependence from the internal relations.
In many cases (but not in all) we can functionally separate the influence
of external and internal relations, so that we can rewrite Eq. (6.1) in the
following form:
dV
= R(t) + f (V) (6.2)
dt
1 As a reminder: First-order means that only the first derivative of V with re-
spect to time appears. Systems with higher derivatives, which are often found
in physics, can be transformed into multi-dimensional first-order systems (see
Example 5.11).
ONE-BOX MODELS 127
dN
= kp (N ) · N (6.3)
dt
For kp (N ) we choose a function that decreases linearly from a max-
imum value kp◦ at N = 0, reaching zero at N = Nmax : kp =
kp◦ (1 − Nmax
N
). Inserting this in Eq. (6.3) yields:
dN N
= kp◦ (1 − )N ≡ f (N ) (6.4)
dt Nmax
The growth rate kp◦ has the dimension [ T ]−1 , Nmax has the same
dimension as N , e.g. a population number, or a population number
per surface area.
For the following considerations it will be helpful to plot the change function
f (N ). In our example, f (N ) is a parabola with a downward facing opening
that intersects the N -axis at N = 0 and N = Nmax (Fig. 6.1a). These two
N -intersects of f (N ) show where the function N (t) “stands still”, that is,
where dN dt = 0. Thus, they are called fixed points. The linear models we dealt
with so far had at most one fixed point, which we called the steady state.2 Ima
fixed point!
Now, however, we have two. This begs the question whether the system is
moving towards one of the fixed points, and if so, towards which one.
In the next section we will discuss a general recipe to answer that ques-
tion. For our current example, it is easy to see that the population N (t)
develops towards the value Nmax except in the case where the initial value
N 0 is zero, since from nothing, nothing can grow! In fact, the differen-
tial equation (6.4) can be analytically integrated, which gives the following
result3 :
N0
N (t) = Nmax ◦ (6.5)
(Nmax − N 0 ) e−kp t + N 0
2 From now on we will only speak of fixed points, since the concept of a
steady state makes less sense as we move towards higher-dimensional and more
abstract models.
3 Since discussing fundamental mathematical principles is not the aim of this
book, we will not go through the integration of Eq. (6.4) in detail. Just one
note: the solution can be derived by using the variable separation method.
128 NONLINEAR MODELS
Change function f (N )
(b) behavior of the popula-
Population N (t)
0.8
tion size N (t) (see Eq. 6.5)
for kp◦ = 1, Nmax = 1 and 0.1 0.6
N 0 = 0.01
0.4
0
0.2
Nmax
-0.1 0
0 0.2 0.4 0.6 0.8 1.0 1.2 0 2 4 6 8 10
Population N Time t
dV
= g(V), (6.6)
dt
We also allow for a possible constant external relation, incorporating it into
the change function g(V), and treat the system as autonomous. We assume
that g(V) has k roots V∞ i (i = 1, . . . , k), that is, the following holds for all
V∞
i :
g(V∞i ) =0 for i = 1, . . . , k (6.7)
The V∞i are then fixed points of the differential equation (6.6). If the system
variable V reaches a fixed point, the model stays put at that position,
because:
dV
=0 (6.8)
dt V∞
i
ONE-BOX MODELS 129
V cannot move away from this point, except if the external relation R (and
with it, g(V)) were to change.
The fixed points of a one-dimensional systems are therefore the roots of
g(V). If g(V) is an n-degree polynomial, there is a maximum of n real roots. Intervals between
These, together with the virtual fixed points V = ∞ und V = −∞, divide fixed points of a one-
the V-axis into (n + 1) intervals (Fig. 6.2a). Because a one-dimensional dimensional nonlinear
system can only move along one variable axis, each fixed point represents a model are called invariant
barrier that the system cannot overcome. Sections between two neighboring areas.
fixed points are called invariant areas. They are a characteristic feature of
one-dimensional models. In two- and n-dimensional models, the system can
move in a two or n-dimensional space (see Fig. 5.2). Therefore, fixed points
can usually be “bypassed”.
a
g( )
Initial state
0 ∞ ∞ 0 ∞
1 2 3
dg
b Line with slope d ∞
G( ) 2
∞ ∞ ∞
1 2 3
Basin of attraction
Fig. 6.2: (a) A one-dimensional system with three fixed points partitions the
variable axis into four so-called invariant areas (I–IV). From the initial state
V0 , the system can only move to the fixed points V∞ ∞
2 and V3 . It cannot leave
the invariant area III, however. (b) Topographic illustration of the same stability
conditions as in (a), as what we call a stability landscape. The ball (which is
infinitely damped, i.e., does not behave like a ball in reality) stays still at V∞ ∞
2 . V1
∞
and V3 are unstable positions, at which a small nudge in either direction will set
the ball in motion. Relation between g(V) and G(V): g(V) = − dG(V) dV
Let’s imagine a model with an initial state V0 lying inside area III of
Fig. 6.2a. So far, we don’t know towards which of the neighboring fixed
130 NONLINEAR MODELS
points this model will move. But we can attempt to answer the question
intuitively: in area III, as we can see in Fig. 6.2a, g(V) is negative. According
to Eq. (6.6), dV/dt < 0, that is, V decreases and the model moves leftwards
on the V axis. This movement keeps up until the model reaches the fixed
point V∞2 , where it stops and remains.
To make these thoughts more formal, we can look at the change function
g(V) in the immediate vicinity of the fixed point V∞ ∞
i , e.g. at Vi + ε with
∞
a small ε. We can then describe g(V) at the point Vi as a Taylor series4
and abort after the linear term:
∞ ∞ dg(V)
g(Vi + ε) ≈ g(Vi ) + ε (6.9)
dV V∞
i
However, since V∞
iis a fixed point, g(V∞
i )
= 0. Thus, it follows from
Eqs. (6.6) and (6.9):
dV ∞ dg
= g(V + ε) ≈ ε (6.10)
dt V∞ +ε i
dV V∞
i i
dV |Vi
dg
Because the derivative is independent of ε, our approximation in
∞
Eq. (6.10) replaces the function g(V) with a straight line in the vicinity of
V∞i . In other words, we have created a linear model out of a nonlinear one.
∞
Figure 6.2a shows the linearized velocity function in the proximity of V2 .
dg
If we define k ≡ dV ∞ 5 and describe ε as the deviation of the momentary
V2
state from the fixed point V∞ ∞
2 , ε = V − V2 then Eq. (6.10) becomes:
dV
= k(V − V∞ ∞
2 ) = −kV2 + kV (6.11)
dt
This is nothing else than the linear inhomogeneous differential equation
(4.5) with constant coefficients. If k < 0, the solution tends to the fixed
point V∞ 6
2 . For k > 0 the model moves away from it. For k = 0 the system
is indifferent: it just stands still where it is. These statements are valid
whether the initial state is to the left or right of the fixed point in question,
as long as the absolute value of the deviation |ε| is not too large. The fixed
point V∞2 in Eq. (6.2) “attracts” the system inside area II as well. We thus
call V∞
2 a stable fixed point or an attractor. Areas II and III together form
its basin of attraction.
In contrast, the model moves away from the fixed point V∞ 3 in both
directions as soon as it deviates to the left or right by the small value ε.
valid for the vicinity of V∞2 , this is of course not generally true for a nonlinear
model.
ONE-BOX MODELS 131
Attractor
Stardate 476.3 - Log of the According to Houston we will fly past 26 years later...
shipwrecked astronauts Dang and Ding... the sun and land on Mars in about 30
years. No problem, we have a base Houston, it s getting hot here! You
there. In the meantime, we can relax... didn t miscalculate the attractor
by any chance!?! HOUSTON!...
Houston, Haha!
do you copy...
Formally, this can also be seen from Eq. (6.11) since here, k > 0. This
is called an unstable fixed point. It is like a needle standing on its tip:
theoretically, there is a point where it can stand still, but in reality even
just the disturbance caused by thermal movements prevents the needle from
remaining upright.
Figure 6.2b shows the stability relations by using a topographic function
G(V), where g(V) = − dG dV . Stable fixed points correspond to a minimum,
unstable ones to a maximum of the topographic function G(V).
The stable and unstable fixed point are shown in Fig. 6.3 (cases 1 and 2).
The figure further shows four functions (cases 3 through 6) whose slope
dg
k ≡ dV ∞ is zero at the fixed point. Cases 3 and 4 are one-sided attractors;
Vi
we call these fixed points one-sided stable. In case 5, within a finite area
around the fixed point, the model does not move. This fixed point is called
indifferent. Finally, case 6 shows a velocity function whose first nonzero
term of the Taylor series is the third differentiation. An example of such a
curve would be g(V∞ 3
6 + ε) = αε . If α < 0 (corresponding to case 6 in the
figure), the fixed point is stable, for α > 0 it is unstable.
Let’s summarize: the slope of the change function g(V) determines the
character of the fixed point, according to the following rules:
dg
<0 Stable fixed point V∞
dV V∞ i
i
dg
>0 Unstable fixed point V∞ (6.12)
dV V∞ i
i
dg Higher-order differentiations of
=0
dV V∞ g(V) need to be considered
i
If a model only has one attractor towards which all systems initiated in Linear models have one
finite space move, that attractor is called asymptotically stable, or global. attractor at most. It is
Since the change function of a linear model is a straight line, a linear model asymptotically stable (or
has one root at most, and thus a single fixed point. This can be stable or global).
132 NONLINEAR MODELS
g( )
∞ ∞ ∞
1 2 3
±ε ±ε ±ε
0
1 2 3
g( )
∞ ∞
4 6
±ε ±ε
∞
5
0
4 5 6
f (V∞
i ) = −R(t) for i = 1, . . . , k (6.13)
The dependence of a sys- We have seen that for a constant external relation Ro , nonlinear models
tem on its history is called can have several possible fixed points V∞ i . If Ro is changed adiabatically
hysteresis. (that is, slow enough), the V∞i change along with it, too. In the process, a
phenomenon called hysteresis can occur.7
We now want to show that hysteresis always occurs when the system
has several fixed points for a given constant value of the external relation
R. As a starting point, we choose Eq. (6.2) and replace R(t) by the constant
value J. According to Eq. (6.13) the fixed points are defined by the following
relation:
f (V∞
i ) = −J (6.15)
If f (V) is a continuous function and intersects the line −J at least twice,
then it must alternately intersect −J from above and below. In the first
case, the slope of f (V) is negative and, according to Eq. (6.12), the fixed
point is stable. In the second case, the fixed point is unstable. Thus, stable
and unstable fixed points alternate.8 Figure 6.5 shows a function f (V) that
has three roots, the first one at zero. One could describe f (V) by the third-
order polynomial
a b ∞
f( ) ∞
3
b
fB B
∞ ∞ ∞ J = − fA
1 A 2 3
a ∞ a
B b 2
−J
J = − fB
fA
A
∞
1
J
dV
= J − kV + ϕ(V) = J + f (V) (6.17)
dt
10 SeeScheffer et al. (2002). The authors give several examples of systems
whose behavior can be described by Eq. (6.17). One of them forms the basis
for Example 6.3, see below.
136 NONLINEAR MODELS
c d
ONE-BOX MODELS 137
In many cases, Eq. (6.17) describes a system for which only positive
values of J and k make sense (see Example 6.3 below). As Fig. 6.6b demon-
strates, this can imply that the jump to the upper fixed point branch is an
irreversible process, because the corresponding jump back would have to
occur under a negative value of J (area marked grey in Fig. 6.6b). If, how-
ever, kVcrit > ϕ◦ (Fig. 6.6c, d), both transitions occur at positive values of
J. In this case, for J = 0 the system only has one fixed point V = 0.
We can use the following simple model to describe the situation described
above. We choose the concentration of suspended phytoplankton B as our
system variable. B is simultaneously a measurement for the turbidity of
the water column:
dB
= JN + f (B) (6.19)
dt
with
−fs (B) for B ≤ Bcrit
f (B) = (6.20)
−αfs (B) for B > Bcrit
138 NONLINEAR MODELS
Phytoplankton elimination
−fs (B) − fs (B)
Bcr it
Phytoplankton/turbidity B
b
s
te
hy
Phytoplankton/turbidity B
op
f1 (B)
r
ac
tm
ou
Critical
tes
ith
hy
W
p turbidity Bcr it
ro
ac
it hm
W
Nutrient concentration JN
c
Portion of the lake surface
with macrophytes
0.4
0.3
0.2
0.1
Fig. 6.7: (a) Elimination function of the turbidity with and without macrophytes.
Note that qualitatively, the function resembles the one in Fig. 6.6c. (b) Steady-
state turbidity as a function of nutrient concentration. (c) Measurements in Dutch
lakes show that hysteresis in the system is indeed related to the distribution of
macrophytes. The macrophytes disappear at a higher phosphorus concentration
than the one where they reappear as the concentration decreases (From Scheffer
et al. (2002), data in (c) from Meijer (2000))
which factor the function fs (B) is reduced if the critical nutrient concen-
tration Bcrit is overshot and the macrophytes die off.
Figure 6.7b shows the fixed-point phytoplankton concentration (and
accordingly the turbidity) B ∞ as a function of nutrient input JN . We get
B ∞ from Eq. (6.19) by setting dB/dt = 0:
f (B ∞ ) = −JN (6.21)
The function ks (C) is plotted in Fig. 6.8. In reality, ks (C) would de-
crease continuously, rather than abruptly as in our model. Eq. (6.23)
has the advantage that the model can be analyzed relatively easily
while retaining all important properties of a realistic model.
140 NONLINEAR MODELS
If we insert Eq. (6.23) into Eq. (6.22), the following model equation re-
sults:
dC ks C for C < Ccrit
= kw Cin − kw C − (6.24)
dt 0 for C ≥ Ccrit
The fixed-point lines are plotted in Fig. 6.9. Line A (C ∞ < Ccrit ) has a
slope of kwk+k
w
s
, line B a slope of 1. The horizontal line at Ccrit delimits the
scope of the two fixed-point lines. We note that for an input concentration
Cin between Ccrit and C = kwk+k w
s
Ccrit > Ccrit , both fixed-point lines are
technically valid. How does the model really behave?
Let’s imagine that initially, Cin is so small that the corresponding fixed-
∞
point concentration is CA < Ccrit . If Cin now grows so slowly that the
Cin
kw +ks
Ccr it C = kw
Ccr it
∞ kw kw kw + ks
CA (C ) = C = Ccrit = Ccrit (6.26)
kw + ks kw + ks kw
At Ccrit , the model now jumps to the upper fixed-point line B. If, on the
other hand, the model moves downwards on B coming from large values of
Cin , it won’t jump back to A at Cin = C but only at Cin = Ccrit .
To summarize: despite its simplicity, the model (6.24) has allowed us to
see two phenomena that are typical for nonlinear models (and do not occur
in linear models):
1. The model has critical states in which even the smallest of changes
in the input (external relation) lead to large changes in the behavior
of the system variable.
12 Question to the reader: How should we specify the attribute slow here?
142 NONLINEAR MODELS
Synergism
...action!
1, 2, and...
Grr.. a nice
synergism!
Synergism: interaction be- To understand more clearly what we mean by “influence each other”
tween different external we remember those systems where a mutual influence doesn’t exist: linear
forces, so that the over- systems. In Sect. 4.3 we looked at a linear model with time-dependent input
all influence is smaller or and saw that the input can be divided into many individual input chunks—
larger than the sum of the the system “processes” each input as if the others didn’t exist (see Fig. 4.5).
individual influences. The combination of all input events is obtained from the sum (superposi-
tion) of all individual events. In general, this method cannot be applied to
nonlinear models.
To illustrate this, we can think of the following question: what is the in-
fluence of water polluted with zinc and mercury on the growth of a specific
algae species? We can assume that each of the two heavy metal concentra-
tions on its own would reduce the algae growth rate by 10 %. If the effects
on algae growth of the two heavy metals were independent of each other,
the overall effect would be a growth rate reduction of 20 %. This is the
case without synergism. In reality, however, we will often observe that the
effects of two pollutants mutually reinforce each other, so the reduction
of growth rate might be more than 20 %. Some substances may also com-
pensate each other’s effects (think poison and antidote). In both cases we
speak of synergism.
Time t
t1 t2
13 Note that here we slightly modify the notation introduced in Eq. (6.7). We
leave out the symbol ∞, the numbering of the fixed point k becomes the upper
index, and the lower index indicates the variable.
144 NONLINEAR MODELS
If the system were linear, it would have at most q = n solutions. For sys-
tems of nonlinear equations, the number of fixed points q can be arbitrarily
large.
To examine the system’s behavior around a fixed point k, we can build
on the method that we used for one-dimensional models in Sect. 6.1.2. Be-
cause it is difficult to imagine surfaces in more than three dimensions, we
first look at a two-dimensional system with the variables V1 and V2 . We
can represent each of the two velocity functions gi (i = 1, 2) as a surface
in three-dimensional space. Two of the dimensions are spanned by the two
variables themselves. This is the phase space as we have already encoun-
tered it, or more precisely, the phase plane. Along the third axis, standing
perpendicular to this plane, we plot the value of g1 or g2 . This way we get
two topographical maps with mountains and valleys. Where both of these
topographies intersect the phase plane there is a fixed point. Around this
fixed point, we can approximate the development of g1 and g2 with tan-
gents in the direction of V1 or V2 , just as we did in Eq. (6.9) (or Fig. 6.2).
∂g1 ∂g1 ∂g2 ∂g2
In total we get four tangents, with the slopes ∂V 1
, ∂V2
, ∂V1
and ∂V2
. The
symbol for the partial derivative (∂) indicates that the slope is being deter-
mined along the chosen direction (V1 or V2 ) and that the other coordinate
is held constant. Applying this principle to more than two dimensions, we
get n2 such slopes. Below, we will explicitly plot these velocity surfaces for
Example 6.6 (Fig. 6.13).
Somewhat more formally, we assume that the system is arbitrarily close
to the fixed point k, so:
Because by definition, at the fixed point itself all change functions gi are
zero, we can approximate them in the vicinity of Vk with the following
Taylor series, aborting after the first-order derivatives:
n
∂gi
gi (V) = gi (Vk + ε) = · εj i = {1, . . . , n} (6.31)
∂Vj
j=1 Vk
The vertical line with the argument Vk tells us to calculate the partial
derivatives at the fixed point Vk . Because Vk is constant in time, dVi /dt =
dεi /dt. Thus, as an approximation, Eq. (6.27) becomes an n-dimensional
linear system of the deviation variables εi :
n
dεi
= Bi,j (Vk )εj i = {1, . . . , n} (6.32)
dt j=1
MULTIDIMENSIONAL MODELS 145
with
k ∂gi
Bi,j (V ) = (6.33)
∂Vj Vk
where the argument Vk indicates that the derivatives of the velocity func-
tions gi in the matrix are to be calculated at the selected fixed point k. Finally something
useful to come
B is the so-called Jacobian matrix. It describes the system of nonlinear out of maths,
equations in an approximated linearized manner in the vicinity of the fixed this Jacobian
mattress!
points. Written in matrix form, Eq. 6.32 is:
dε
= B(Vk )ε (6.35)
dt
The expressions (6.32) and (6.35) are thus linear approximations of the
nonlinear system (6.27). By way of the eigenvalues of the Jacobian matrix
we can now analyze the behavior of the model around the fixed point. We
must keep in mind however that under some circumstances such a stability
analysis may only be valid in very small ε surroundings of Vk . Furthermore,
there are nonlinear models whose velocity functions gi are not differentiable,
so that no Jacobian matrix can be calculated. Finally, there are cases where
the linearized model cannot make any statement about the real model (see
Sect. 6.2.4).
a) b)
c) d)
e) f)
Real eigenvalues
(a) Both eigenvalues are real and negative (λ1 , λ2 < 0): stable fixed point,
also called stable star (Fig. 6.11a).
(b) Both eigenvalues are real and positive (λ1 , λ2 > 0): unstable fixed
point, also called unstable star (Fig. 6.11b).
(c) Eigenvalues are real and have different signs, e.g. (λ1 > 0, λ2 < 0):
Unstable fixed point, also called saddle point (Fig. 6.11c).
Non-real eigenvalues
If the eigenvalues are non-real, λ1 and λ2 only differ in the sign of their
imaginary parts.14 The system variables oscillate around the fixed point in
three possible ways:
(d) Re(λ1 , λ2 ) < 0: The model moves on a spiral towards the fixed point
(Fig. 6.11d); the fixed point is stable.
(e) Re(λ1 , λ2 ) > 0: The model moves away from the fixed point on a
spiral (Fig. 6.11e); the fixed point is unstable.
It can be shown that, except for case f (center), examining the fixed points
by linearization does indeed yield an accurate description of the nonlinear
system.15 This holds at least in a (possibly very small) area around the
fixed point. In the case of a center (case f), we cannot decide whether If the fixed point is a
the behavior of the real (nonlinear) model follows categories d, e or f. We center, the system’s be-
will illustrate this with Example 6.9. Furthermore, there is still another havior in its vicinity
possibility: the convergence of the trajectory towards a closed curve around cannot be determined by
the fixed point. This case is called a limit cycle and we will illustrate it in linearization.
Example 6.8 below.
For nonlinear models with more than two dimensions, the eigenval-
ues of the Jacobian matrix are still useful to analyze the behavior near
the fixed point. If n = 4, for example, the eigenvalues can be calculated
from the solution of a fourth-order equation. The more dimensions a model
has, the more diverse the possible combinations of real and non-real parts
of the eigenvalues become. Thus the diversity of the behavior of fixed
points also rises. In certain cases, e.g. if all solutions are real, we can still
14 Because they are the solutions of a system of quadratic equations, non-real
eigenvalues only occur as conjugate-complex pairs (see Eq. 5.10).
15 As a reminder: the calculation of the Jacobian matrix presupposes that the
k1 X k3 XY
Fig. 6.12: Model by Lotka
k2 Y
Prey X Predator Y and Volterra for the inter-
action between predator Y
and prey X
Because of the two terms ±k3 XY , the model is nonlinear. It has two
fixed points, first, the trivial fixed point X1∞ = 0, Y1∞ = 0, and second:
k2 k1
X2∞ = , Y2∞ = (6.37)
k3 k3
We get this result by setting the left sides of Eqs. (6.36) equal to zero.
To determine the stability of the fixed points, we will analyze the model’s
behavior in the vicinity of the second fixed point by introducing new vari-
ables. This is somewhat more laborious than the method with the Jacobian
matrix, but also easier to understand. We define δX and δY as the deviation
of the population sizes from the fixed point:
δX = X − X2∞ , δY = Y − Y2∞ (6.38)
Since X2∞ , Y2∞ are constant, it follows that:
dδX dX dδY dY
= and = (6.39)
dt dt dt dt
In Eqs. (6.36), X and Y can be replaced by the new variables δX and δY .
After some algebraic conversions, we get:
dδX
= −k2 δY − k3 δY δX
dt (6.40)
dδY
= k1 δX + k3 δY δX
dt
150 NONLINEAR MODELS
250 250
Y∞
2
Prey X, Predator Y
200
X(t) 200
Prey X
150 Y(t) 150
X∞
100 X∞
2
2 100
Y∞
50 2 50
0 0
0 10 20 30 40 0 50 100 150 200
Time t [yr] Predator Y
moves away from it, or it circles it on a closed trajectory like the undamped
harmonic oscillator (see Fig. 5.10).
It can be shown that the Lotka-Volterra model belongs to the third
category (see e.g. Arrowsmith and Place 1992). In Fig. 6.14, X and Y are
shown both as functions of time and in the two-dimensional phase space.
The plot shows that the trajectory describes a closed curve, but not an
ellipse as in Fig. 5.10. Correspondingly, X(t) and Y (t) are also not pure sine
or cosine functions. This non-harmonic (yet undamped) behavior reminds
us that the real system of Eqs. (6.36) is nonlinear and that Eq. (6.41) is
merely a linear approximation.
The kinship between the Lotka-Volterra model and the model of the
harmonic oscillator suggests that they also share the feature of structural
instability. The following example demonstrates that this is indeed the case.
Predator
Example 6.7: Predator-prey model with self-interaction with self-
interaction!
We will now make a minor change to the predator-prey model from
Example 6.6 by assuming that for large prey populations, the prey
animals themselves impede and potentially even kill or eat one another.
We can describe the resulting negative effect on the prey population
dynamics in Eq. (6.36) with an additional term of the form (−k4 X 2 ). k4
is an arbitrarily small coefficient of self-interaction. Note that because
the variable X appears squared, the new term is only relevant for large
values of X. The modified system equations are:
dX
= k1 X − k3 XY − k4 X 2
dt (6.42)
dY
= −k2 Y + k3 XY
dt
152 NONLINEAR MODELS
The system still has two fixed points. The trivial fixed point (X1∞ =
Y1∞ = 0) remains unchanged. The nontrivial fixed point however is slightly
changed compared to Eq. (6.37):
k2 k1 k2
X2∞ = , Y2∞ = − ; with = k4 (6.43)
k3 k3 k3 k3
Problem 6.4 asks you to calculate the Jacobian matrix of Eq. (6.42) and
determine the eigenvalues at the nontrivial fixed point. From the results
of that calculation, the deviation of X and Y from the nontrivial fixed
point is described by the following approximative system of linear equa-
tions:
dδX
= −k2 δY − δX
dt (6.44)
dδY
= (k1 − )δX
dt
We assume that the prey’s self-interaction is arbitrarily small but nonzero.
In particular, k1 holds. The system (6.44) thus essentially differs from
the original equations (6.41) only by the additional term −δX in the first
equation (just like the damped pendulum, Eq. 5.63). We already know what
this means for the eigenvalues (see Eq. 5.67)—they receive a (very small)
negative real part:
λi ≈ − ± iω, ω = k1 k2 (6.45)
2
With that, the fixed point (X2∞ , Y2∞ ) becomes stable, that is, an at-
tractor (Fig. 6.11d). The oscillation of the predator and prey populations
is damped around the fixed point. The system eventually assumes the con-
stant values X2∞ and Y2∞ . Furthermore, because the fixed point is no longer
a center (as it was in our first predator-prey model), we know now that the
nonlinear system qualitatively behaves like its linearized counterpart. Fig-
ure 6.15 shows the dynamic behavior of the modified predator-prey model.
The Lotka-Volterra model The Lotka-Volterra model profoundly influenced the thinking of the first
profoundly influenced the ecological modelers. The Hudson Bay Company’s 100-year catch statistics
beginning of ecological sys- of the Canada lynx and its prey, the snowshoe hare, were often cited as
tems modeling. a textbook example of the periodic oscillations of a predator-prey pair as
predicted by the model (Fig. 6.16). The catch statistics show the population
density of both species. It exhibits a regular fluctuation with a period of
about 10 years. With some goodwill, we can see that the maximum hare
population does indeed occur just before the maximum of the lynx—as
predicted by the Lotka-Volterra model.
This apparent confirmation of the Lotka-Volterra model does not stand
up to critical examination, however. First, we know from our mathematical
considerations that the model is structurally unstable. Thus there are many
reasons why the fluctuations should disappear. Second, the snowshoe hare
population also fluctuates in areas where there are no lynxes. And finally,
the lynx does not subsist exclusively on a snowshoe hare diet.
MULTIDIMENSIONAL MODELS 153
200 200
Prey X
Y(t) X2
¥
100 X2 100
¥
50 Y2 50
Y2
0 0
0 10 20 30 40 0 50 100 150 200
Time t [yr] Predator Y
Fig. 6.15: Modified predator-prey model: compared to Fig. 6.14, the dynamic prey
equation has an additional prey-prey self-interaction term (Eq. 6.42). Therefore
the oscillations of the predator and prey populations Y and X are damped and
the system moves towards the nontrivial fixed point. The position of the fixed
point itself and the period of the oscillations change only very slightly for a small
self-interaction. Parameters and initial condition are like in Fig. 6.14, and k4 =
5 × 10−4 a−1
160
140 Hare
Lynx
120
Number in thousand
100
80
60
40
20
1845 1855 1865 1875 1885 1895 1905 1915 1925 1935
Fig. 6.16: Population fluctuations of snowshoe hare and lynx, from the catch
statistics of the Hudson Bay Company. This is a classic example of cyclical os-
cillation of population densities (Source: MacLulich 1937 cited in Odum 1971)
In Eq. (6.46), the first term describes the logistic growth of the prey
(see Example 6.1). Xk is the hypothetical equilibrium population of X
if there were no predators. The second term of Eq. (6.46) simulates the
decimation of prey by the predators. In contrast to the Lotka-Volterra
model, the predators have a limited appetite: for X KX , it reaches
the value −wY which is independent of X. Eq. (6.47) describes the
growth of the predator population. It is only positive if JY /X < 1, i.e.
if Y < X/J. In other words: the predator population can only grow if
at least J prey animals are available per predator.
250 250
Fig. 6.17: Numerical
Y¥
Prey X, Predator Y
200 200
solution of the Holling-
X(t) Tanner model. The sys-
150 150 tem moves towards the
Prey X
limit cycle. Parameters
X¥
100 X¥ 100 used: r = 2.5 year−1 , s =
Y(t) 0.225 year−1 , w = 5 year−1 ,
50 Y¥ 50 Xk = 300, KX = 50,
J=2. Initial conditions:
0
0 10 20 30 40 50 20 30 40 50 60
0 X 0 = 50, Y 0 = 60
Time t [yr] Predator Y
Model 1:
dx
= −y + x(x2 + y 2 )
dt (6.50)
dy
= x + y(x2 + y 2 )
dt
Model 2:
dx
= −y − x(x2 + y 2 )
dt (6.51)
dy
= x − y(x2 + y 2 )
dt
The only difference between the two models is the sign of the second term
on the right-hand side of the equations. Both models have the fixed point
x = y = 0. The Jacobian matrices at the fixed point are identical:
0 −1
B= (6.52)
1 0
We already know that the eigenvalues are purely imaginary (λi = ±i).
The fixed points of the linearized systems are centers. But how do the
nonlinearized systems really behave in the proximity of the fixed point? To
answer this question, we introduce two new variables r and θ to replace x
and y: y
r = x2 + y 2 , θ = arctan (6.53)
x
Of course, we didn’t choose these particular notations for r and θ by ac-
cident: they are the polar coordinates in the original phase space (x, y).
Now we have to formulate the new dynamic equations, which requires
Center
Test series #1 complete:
Center is an attractor.
MULTIDIMENSIONAL MODELS 157
some effort, but nothing more than the chain rule of differentiation. We
demonstrate the procedure with model 117 :
dr dr dx dr dy
= +
dt dx dt dy dt
x dx y dy
= +
x2 + y 2 dt x2 + y 2 dt
−xy xy
= + x2 x2 + y 2 + + y 2 x2 + y 2
2
x +y 2 2
x +y 2
3
= (x2 + y 2 ) 2 = r3
(6.54)
dθ dθ dx dθ dy
= +
dt dx dt dy dt
−y/x2 1/x
= 2
[−y + x(x2 + y 2 ] + [x + y(x2 + y 2 ]
1 + (x/y) 1 + (x/y)2
y2 x2
= 2 − xy + + xy = 1
x + y2 x2 + y 2
For model 2, the result is almost identical, except that dr/dt has an inverted
sign:
dr dθ
= −r3 ; =1 (6.55)
dt dt
The new variables r and θ have decoupled the two systems of differential
equations (6.50) and (6.51). The arc function θ(t) has the trivial form:
θ(t) = θ0 + t (6.56)
This means that in both cases, the trajectory turns with constant angu-
lar velocity in a (mathematically) positive direction of rotation (which is
counter-clockwise) around the coordinate origin. The difference between
the two models is that in the case of Eq. (6.50) (model 1) the radius grows
with time (outward spiral), whereas in the second case, the radius decreases
(inward spiral) until the system comes to a rest in the fixed point x = y = 0.
The two solutions are shown in Fig. 6.19 together with the trajectory of the
undamped linear oscillator (r = const). Note that the third graph, the lin-
ear approximation of both nonlinear models, is not capable of making a
distinction between them.
17 Remember that d
dz
(arctan z) = 1
1+z 2
.
158 NONLINEAR MODELS
chaotic if two trajectories starting from arbitrarily close initial points sud-
denly develop completely differently. Often, chaotic systems remain close
to an attractor for a certain timespan, only to suddenly jump into a com-
pletely different course that unfolds around another attractor. Such at-
tractors are called strange attractors. The chaos is called deterministic (as
opposed to stochastic) because the underlying equations are entirely de-
terministic, i.e., they contain no random quantities. Because in reality, a
system’s state can never be measured to perfect precision (no measurement
is without measurement error), there are always tiny differences between
apparently identical states. These differences can be responsible for the
diverging development from two seemingly equivalent initial states.
A well-founded discussion of deterministically chaotic systems would go
beyond the scope of this book. We merely want to discuss one example:
the Lorenz system. In order for deterministic chaos to occur, the system of
differential equations has to be at least three-dimensional.18
40 40
35 35
30 30
z 25 25 z
20 20
15 15
10 10
−20 −10 0 10 20 −20 −10 0 10 20
x y
Fig. 6.20: Solution of the Lorenz model (Example 6.10) with the parameters
a = 10, b = 28 and c = 8/3. Initial conditions (x0 , y 0 , z 0 ) = (−11.3, −13.3, 28.0).
Attractor at (±8.4853, ±8.4853, 27.0000). On the left, trajectories in three-
dimensional phase space are projected onto the (x, z) plane, on the right, onto
the (y, z) plane
(1) x = y = z = 0
(2) x = y = (b − 1)c, z = b − 1
(3) x = y = − (b − 1)c, z = b − 1
Question 6.3: Give examples of systems where the implicit and explicit
time dependence cannot be broken up as we did in Eq. (6.2). You can modify
systems we introduced previously to get such examples.
Question 6.4: Give some properties that distinguish nonlinear from linear
systems.
ination rate is k = 0.005 d−1 for the first 6 months of the year, while for
the rest of the year it is k = 0.015 d−1 . Analyze the resulting model, in
particular by answering the following questions:
(a) Establish the model equation and determine the fixed point(s) and
its/their stability.
(b) Draw the model space graphically with the aid of the individual com-
ponents of the change function (as we did in Fig. 6.4).
(d) Calculate the maximum allowable specific rate kf so that the fish
population does not go extinct.
(b) What are the conditions for the coefficients ki so that a nontrivial19
fixed point exists?
(c) What does the condition found in (b) mean biologically? How could
we meaningfully modify the model (i.e. by introducing an improve-
ment which makes sense from a biological point of view) so that a
nontrivial fixed point always exists?
dB
= (k3 − k4 B)B if A=0
dt
(c) Determine the stability properties of those fixed points where both
populations are different from zero. Hint: use the Jacobian matrix.
Use the following values for question (c) and assume that A and B are
non-dimensional:
Time-Discrete Models
As a starting point we choose Q(0) = 0. The relation in Eq. (7.1) can either
be proven mathematically or simply checked for its validity by explicit
calculation of Q(1) , Q(2) and so on. We thus have a recipe to determine
Q(n+1) from Q(n) , that is, how to calculate (26)2 if we already know that
(25)2 = 625: (26)2 = 625 + 2 × 25 + 1 = 676. If we interpret Q(n) as a time
In discrete models, differ- We then write Eq. (4.7) for a finite time step with a so-called difference
ence equations take the equation:
place of differential equa- dV ΔV V(n+1) − V(n)
tions.
−→ = = k · V(n) (7.3)
dt Δt Δt
Solving for V(n+1) gives us:
If we begin with the initial state V(0) and apply the equation m consecutive
times, we get:
V(m) = (1 + Δt · k )m V(0) (7.5)
This expression does not seem to have much in common with the exact
solution of the differential equation (4.7), which, as we learned in Eq. (4.8),
is an exponential function. Of course, the numerical approximation of the
solution of (4.7) should eventually transform into the exact solution if only
we choose the time step Δt small enough. If the integration should take
place over the entire time period T , a given time step results in m = T /Δt
iterations. In the limit case, we have Δt −→ 0 and m −→ ∞. Inserted in
Eq. (7.5), V(m) ≡ V(T ) thus becomes a limit of the form:
k T m (0)
V(T ) = lim [(1 + ) ]V (7.6)
m→∞ m
A m
As can be learnt from mathematical handbooks, limm→∞ (1 + m) =
eA , so with A ≡ k T :
V(T ) = ek T V0 (7.7)
This result is in full accordance with Eq. (4.8).
Let us exemplify the relation between discrete and continuous temporal
change with an example which is familiar to all of us: the calculation of
interest on a bank account.
(b) N (t = 14 ) = N0 (1 + p4 ), thus
N (t = 24 ) = N (t = 14 )(1 + p4 ) = N0 (1 + p4 )2
...
⇒ N (t = 1) = N0 (1 + p4 )4 = (1 + 0.05 4
4 ) N0 = 1050.95e
Fortunately for us depositors, the fact that interest is only payed once a
year does not make such a big difference for small interest rates. For larger
rates, it would make a difference however. For instance, for a rate of 20 %
per year the balance would grow by 22.14 % over a year for continuous
interest payment, compared to the “mere” 20 % if the interest is added
once per year.
In many (but not in all) cases, the numerical solution of a differen-
tial equation approaches the exact solution if the time step is chosen small
168 TIME-DISCRETE MODELS
3 We don’t want to suggest that the “forward integration” we used in Eq. (7.3)
is a particularly good strategy for numerical integration. There are in fact more
efficient and reliable methods.
TIME-DISCRETE MODELS WITH ONE VARIABLE 169
(a) Calculate the fish population N (n) at the beginning of 2001 and
2002.
(b) In the long term, what is the fish population at the beginning of
the year?
(c) If, at the beginning of 2000, there had been no fish in the
pond, how long would it take for the difference to the first case
(N 0 =1,000 in the year 2000) to be less than ten fish?
n
1
oscillating convergence, 0.4
(c) a0 > 1: direct diver-
0.5
gence, (d) a0 < −1: oscil- 0.2
lating divergence
0 0
c d
300 50
250
200 0
n
150
100 50
50 a0 = 2 a 0 = −2
0 100
0 2 4 6 8 0 2 4 6 8
n n
(a) We begin our analysis in the year 2000 (N 0 =1,000) and apply
Eq. (7.8) with a0 = 0.4, I =1,200:
In the year 2001: N (1) =1,200+0.4×1,000 = 1,600
In the year 2002: N (2) =1,200+0.4×1,600 = 1,840
(b) Because |a0 | < 1, in the long run, at the end of each winter (i.e.
just before the new fish are set free) the fish population reaches the
following value (Eq. 7.14):
1, 200
N∞ = − = 2, 000
0.4 − 1
I
N (n) = (1 − an0 ) + an0 N 0 ,
(1 − a0 )
we can see that the influence of the initial state N 0 is limited to the
last term. In the first case, this term is (0.4)n × 1,000, in the second
case it is zero. Thus we must ask ourselves for which n the following
relation is valid:
(0.4)n × 1, 000 = 10
TIME-DISCRETE MODELS WITH ONE VARIABLE 171
"
q
If ai = 0, and if we choose
i=0
I
b= (7.17)
"
q
ai
i=0
then an iterative equation for V̂(n) emerges which, apart from the miss-
ing term I, has exactly the form of Eq. (7.15). Note that the solution of
Eq. (7.15) requires the specification of q initial conditions (e.g. V(0) , V(1) ,
. . ., V(q−1) ).
5 Either the natural logarithm or the common (base-10) logarithm work. The
following numbers correspond to the common logarithm.
172 TIME-DISCRETE MODELS
The sale of the product begins in month 1 with the first 10, 000 bottles.
N (0) = 0
N (1) = 10, 000
Without having to calculate much, we can easily answer the question about
the long-term availability of bottles (that is, the fixed point N ∞ ). Twenty
percent of the bottles never come back, which means they need to be
replaced by new bottles (10,000 per month). This implies that, at the fixed
point, 50, 000 bottles are filled each month, since 20 % of that are the 10,000
new ones. The bottle inventory develops as follows:
At the fixed point, the following must hold: N (n+2) = N (n+1) = N (n) =
N ∞ . Therefore:
Calculating the monthly values N (n) with the “recipe” in Eq. (7.18) is easily
done on a computer; this would show that the value indeed asymptotically
approaches N ∞ = 50, 000.
TIME-DISCRETE MODELS WITH ONE VARIABLE 173
Recycling bottles
Boohooo, stop! I
Answer me,
don t know!
professor!
Sigh..
In Sect. 7.1.2 we saw that a time-discrete model can also lead to divergent
solutions (see Fig. 7.1). It would therefore be useful to have a mathematical
tool to analyze equations of the form (7.15) more generally.
To that end, we look at the homogeneous form of Eq. (7.15), which we get
via the transformation in Eq. (7.16):
V̂(n) = λn (7.20)
a0 λn+q + a1 λn+q−1 + . . . + aq λn = 0
1. The absolute value of all λj is smaller than 1: |λj | < 1. Then, V̂(n)
converges to zero for n → ∞.
6 Equation (7.20) makes it clear why we prudently write V(n) : The n in paren-
theses gives us the n-th iteration step, whereas λn actually means the n-th
power.
174 TIME-DISCRETE MODELS
Obviously, the roots of the function g(V(n) ) are the fixed points of the
model, since there, according to (7.26), V(n+1) = V(n) . As in the case of
nonlinear differential equations we analyze the behavior of the system at
the fixed point V∞
j by introducing a new discrete variable:
X(n) = V(n) − V∞
j (7.28)
If the deviation from the fixed point (X(n) ) is small, g can be developed
into a Taylor series and broken off after the linear term (see Fig. 7.2):
∞ ∞ dg
g(Vj + X ) = g(Vj ) +
(n)
X(n) + . . . (7.30)
dV V∞
j
Since V∞ ∞
j is a fixed point, by definition, g(Vj ) = 0. We get a linear differ-
ence equation for X (n)7
:
dg
X (n+1)
−X =
(n)
X(n) (7.31)
dV V∞
j
7 Obviously, the procedure is completely analogous to the one we used for non-
linear differential equations in Chap. 6.
176 TIME-DISCRETE MODELS
From our earlier discussion (e.g. in Fig. 7.1), we already know that Eq. (7.32)
tends to the fixed point if the following holds:
dg
|A| = 1 + <1 (7.33)
dV ∞
Vj
dg
Expressed by dV this means, in accordance with the continuous case
V∞
j
(Eq. 6.12), that the fixed point V∞
j is unstable if the following holds:
dg
≥ 0 ⇔ V∞ unstable (7.34)
dV V∞ j
j
The inverse is not true however: in fact, because of Eq. (7.33) the fixed
point can also be unstable due to a strongly negative slope of the function
g. Altogether, the conditions for stability are:
dg
0> > −2 ⇔ V∞ stable (7.35)
dV ∞ Vj
j
dg
= −2kV(n) + kb , (7.36)
dV
For the two fixed points this results in:
dg dg
= +kb , = −kb , k, b > 0 (7.37)
dV V∞ dV V∞
1 2
For the following analysis of the logistic growth model we closely follow
Arrowsmith and Place (1992). With the new variable x(n) = 1+kbk
V(n) the
model equation takes the form
x(n+1) = μ(1 − x(n) )x(n) ≡ Fμ (x(n) ) , μ ≡ 1 + kb (7.38)
TIME-DISCRETE MODELS WITH ONE VARIABLE 177
μ−1
x∞ = (7.40)
μ
a b y
y
Fixed point:
Fµ (x)≡µx (1−x)=x
B B’
y=µx(1−x)
Fµ (x (0)) A A’ C
C’
y=x
x x
x∞ 1 x (0) x (1) x
∞ (2)
x 1
Fig. 7.3: (a) Graphical determination of the fixed point of the discrete model
x(n+1) = μx(n) (1 − x(n) ) ≡ Fμ (x(n) ). See text for further explanations. (b) The
movement of the model from the initial state x(0) to the fixed point x∞ can be
determined graphically: After the first iteration, Fμ (x(0) ) takes on the new x value
x(1) , at point A. That x value can be found by mirroring x(1) = Fμ (x(0) ) on the
line x = y (point A’). The value Fμ (x(1) ), belonging to x(1) , lies at point B. By
mirroring B on the line x = y (point B’) we get x(2) or point C, and so forth, all
the way to the fixed point x∞ (From Arrowsmith and Place (1992))
As shown above, the nontrivial fixed point is stable for 0 < kb < 2
which translates into the stability condition 1 < μ < 3. Yet, for μ ≥ 3
there are fixed points as well, although of a different kind. For instance, for
μ = 3.2, the system ends on a closed path that oscillates between the two
178 TIME-DISCRETE MODELS
values x∞ ∞
1 and x2 lying on either side of x
∞
(Fig. 7.4).8 In other words: if
(n)
we only consider every other value x , i.e. only either the even or odd n,
these sequences again reach a fixed point. We thus have to formulate a new
specification from the iterative equation Eq. (7.38) that directly takes us
from x(n) to x(n+2) , skipping the intermediate step x(n+1) . We name this
(2) (2)
“double-jump function” Fμ . To get Fμ out of Fμ , we insert y = μx(1−x)
in the function Fμ (y) = μx(1 − y):
(2)
Fμ is a fourth-order polynomial and must not be confused with the square
of Fμ . Its fixed points can be calculated—as in the case of Fμ —from the
following equation:
x∞ = Fμ(2) (x∞ ) (7.42)
(2)
Graphically, they can be found from the intersection of y = Fμ (x)
with the line y = x.
(2)
In Fig. 7.5 we plotted Fμ (x) for three different values of μ. We are
interested in the nontrivial (x = 0) intersections of the curve with the line
x = y. For μ = 2.8 (Fig. 7.5a) there is a single intersection point x∞ . Of
course, it is identical with the fixed point of the “single-jump function”
Fμ (as per Eq. 7.40, x∞ = 1.82.8 = 0.643), since if the system doesn’t move
anymore after a single jump, it will also remain motionless after two, three,
or q jumps.
(2)
For μ = 3, the curve y = Fμ (x) has a slope of exactly 1 at the
intersection—as does the function y = x (Fig. 7.5b). For even larger values
of μ (e.g. for μ = 3.2, see Fig. 7.5c), the original intersection point x∞ is
divided into two new intersection points x∞ ∞
1 and x2 . These two drift apart
8 Strictly speaking this is only true if the initial point does not lie at x∞ =
μ−1
μ
= (3.2−1)
3.2
= 0.6875.
TIME-DISCRETE MODELS WITH ONE VARIABLE 179
a b c
y=x
y = Fµ(2)(x)
Fig. 7.5: (a) The fixed points of the “double-jump function” (Eq. 7.32) result
(2)
from the intersection of the curves y = Fμ (x) and y = x. (b) For μ = 3,
the system shows a bifurcation: The intersect x∞ splits into the two points x∞
1
and x∞2 (period doubling). (c) “Two-jumps function” for the situation shown in
Fig. 7.4 (From Arrowsmith and Place (1992))
further for an increasing μ. The “double-jump system” then has three fixed
points in total (the middle one, x∞ , being unstable). The “single-jump
system” in contrast jumps back and forth between x∞ ∞
1 and x2 . In fact,
Fig. 7.5c represents the case shown as “single-jump system” in Fig. 7.4.
The behavior of the system at μ = 3 is called phase doubling. The points
at which this doubling occurs are called bifurcation points. In order to get
further bifurcation points, we have to examine the “four-jumps function”
(4) (8) (16)
Fμ , or Fμ , Fμ and so forth. In fact, the model experiences further
phase doublings with growing μ. Yet the story does not end here: Above
the critical value μ = 3.8284 . . ., completely new situations for which no
periodic cycles exist appear on the scene. For μ > 4 these aperiodic fluctu-
ations play the predominant role if the system did not start on a periodic
solution by chance. This strange behavior is another example of determin-
istic chaos, which we already encountered in Sect. 6.2.5.
The emergence of deterministic, but unpredictable (that is, apparently Even one-dimensional non-
chaotic) behavior is not actually that surprising—we know it very well from linear difference models
our own experience. We can take the 104 cards from a Solitaire card game can exhibit chaotic (not
as an example. They can be mixed in 104! ways (factorial of 104, a number easily predictable) behavior.
Bifurcation
A father sends forth his two sons to Nothing but chance brought luck
try their luck in the world. They arrive to one of them while setting the
at a fork in the road and throw a coin to other on the road to ruin.
decide who goes left and who right. The
first son finds a princess and marries
her, the second encounters a dragon
and is eaten... What did
Einstein say? God
does not play dice.
180 TIME-DISCRETE MODELS
with more than 160 digits). If we now deal a certain set of cards from our
deck, and make our decisions strictly according to certain rules, we will be
able to win the game for certain sets of cards but not for others. Switching
two cards could be enough to turn a set without a solution into one with
a solution. Even though all steps are completely determined, we will be
unable to deduce any regularity that could let us predict whether the game
will work out or not.
The pij are the elements of the coefficient matrix P = (pij ) which we
already introduced for multi-dimensional linear differential equations (see
Eq. 5.81). So we can also write Eq. (7.43) in matrix form:
the introduction of new variables, we could use all the tools we learnt for
one-dimensional systems in Sect. 7.1.2 to solve Eq. (7.43).
Instead of more theoretical considerations and the risk of loosing track
of the essential issues, we now want to turn to the concrete example of a
two-dimensional system and use it to illustrate the important concepts step
by step.
First, we establish the iterative equation by translating the rules into math-
ematical expressions step by step. The change of balance in the k-th step
is:
2 1
X (k) − X (k−1) = X (k−1) − Y (k−1)
3 3 (7.46)
2 1
Y (k) − Y (k−1) = − X (k−1) + Y (k−1)
3 3
Eq. (7.46) can be transformed to:
5 (k−1) 1 (k−1)
X (k) = X − Y
3 3 (7.47)
2 (k−1) 4 (k−1)
Y (k) =− X + Y
3 3
182 TIME-DISCRETE MODELS
Whether the bank goes bust or not can be decided by looking at the de-
velopment of the sum S (k) = X (k) + Y (k) throughout the game. If we add
up the two equations from Eq. (7.47), we get:
So there is no danger for the bank; the sum of balances stays constant at
S 0 = X 0 + Y 0 =1,500. On the other hand, however, the bank also doesn’t
make any profit. But what about the distribution of the total amount
amongst the two players?
Table 7.1: Example 7.5: State of the game after the first four rounds
Table 7.1 shows the state of the game after the first four rounds, calculated
by applying Eq. (7.47) successively. Not a good game for Y! Even though
he began with a head start, after only four rounds he’s already in debt.
How does the game go on?
It would be elegant if we could formulate the solution of the system
of equations similarly as the one for the one-dimensional system (Eq. 7.8).
Let’s recall Eq. (7.49): the difference equation of the sum S (k) ≡ X (k) +Y (k)
is solvable by itself, without calculating the individual solutions of X (k)
and Y (k) . We will apply a trick (the benefit of which will become apparent
soon): we’ll look for a second linear combination9 of X (k) and Y (k) whose
iterative equation is solvable by itself. Indeed, such a form exists:
If, with the aid of the original iterative equations (7.47), we replace the
variables X (k) and Y (k) in Eq. (7.50) with their “predecessors” X (k−1) and
Y (k−1) , it follows that:
5 (k−1) 1 (k−1) 2 (k−1) 4 (k−1)
T (k)
=2 X − Y − − X + Y
3 3 3 3
10 2 2 4
= + X (k−1) − + Y (k−1)
3 3 3 3
= 4X (k−1) − 2Y (k−1) = 2 (2X (k−1) − Y (k−1) )
So:
T (k) = 2T (k−1) (7.51)
The solution of the iterative equation now becomes very simple:
T (k) = 2k T 0 = 2k (2X 0 − Y 0 ) = 2k × 300 (7.52)
What remains is to replace the original variables X (k) and Y (k) with the
new variables S (k) and T (k) , whose solutions we know (see Eqs. 7.49 and
7.52). After a bit of algebra we find:
1
X (k) = (S (k) + T (k) )
3 (7.53)
1
Y (k) = (2S (k) − T (k) )
3
By using the solutions for S (k) and T (k) we can thus also reconstruct those
for X (k) und Y (k) :
1
X (k) = (S (0) + 2k T (0) )
3
1
= (1, 500 + 2k × 300)
3 (7.54)
1
Y (k)
= (2S (0) − 2k T (0) )
3
1
= (3, 000 − 2k × 300)
3
It now immediately becomes clear that player Y must loose. In Problem 7.6
we ask how large the initial sums X (0) and Y (0) have to be so that (with
the same rules) player Y wins or the amounts of the two players remain
constant. Of course, the latter situation represents a fixed point.
How could one have guessed the trick we applied to solve Eq. (7.46)?
In the following section, we will discuss the deeper meaning of the solution
and the method we used to get to it.
(Eqs. 7.49 and 7.50). Although it was possible, with some experience, to
see from the system of equations (7.46) or (7.47) that the sum of the two
variables X (k) and Y (k) remains constant, the choice of Eq. (7.50) appears
quite random.
Of course, behind this method lurks a mathematical concept that we
frequently mentioned: the determination of the eigenvalues (and associated
eigenfunctions) of a linear system. In Appendix D we explain the procedure
for a n-dimensional system. Without using computational methods, only
two-dimensional systems can be solved by hand with reasonable effort.
Table 7.2 on the next page summarizes the approach for Example 7.5.
There would be a lot more to say about time-discrete models and their
difference equations (we haven’t even mentioned multi-dimensional non-
linear models in passing!). Nevertheless, we want to close the discussion
here. There is other literature available to delve deeper into the topic (e.g.
Luenberger 1979).
TIME-DISCRETE MODELS WITH SEVERAL VARIABLES 185
with
5 4
tr(P) = + =3
3 3
1
det(P) = (20 − 2) = 2
9
Eigenvalues (D.13):
1 1
λi = [3 ± 32 − 4 × 2] = [3 ± 1]
2 2
λ1 = 2, λ2 = 1
Eigenfunctions (D.17):
2 2 1
z1 = − X + (2 − 5/3)Y = − X + Y
3 3 3
1
= − (2X − Y )
3
1 1
z2 = (1 − 4/3)X − Y = − (X + Y )
3 3
Note: Except for the constant factors, z1 and z2 are consistent with the
variables S und T , that is, z1 = − T3 and z2 = − S3 . The eigenfunctions are
only determined to a constant factor.
Determinant (D.21):
4 5 1 1
D = −2 × 2 + 2 × + 1 × = [−12 + 8 + 5] =
3 3 3 3
So, according to (D.20) the reverse transformation is:
1 5
X = 3 − z1 − (2 − )z2 = −(z1 + z2 )
3 3
4 2
Y = 3 −(1 − )z1 − z2 = z1 − 2z2
3 3
The result conforms with Eq. (7.53), if—as mentioned before—we take into
account that z1 = −T /3 and z2 = −S/3.
186 TIME-DISCRETE MODELS
Question 7.2: Explain the relationship between the terms difference model
and time-discrete model.
Question 7.3: Why is it important whether the interest is paid out yearly
or quarterly for interest rates above 10 % per annum, but much less so for
2 %?
Question 7.9:
(b) What is the answer to that question for nonlinear systems of differ-
ential equations?
After 1 week of marathon training. After 2 weeks of marathon training. After 3 weeks of marathon training.
t
an
mix
v
system is spatially com-
le
re
pletely mixed (homoge-
e
Mixing time
ar
neous). For τr τmix ,
n
io
at
all system components are
rm
r mix
fo
Rapid mixing locally at equilibrium but
ns
tra
d
(spatial inhomogeneities spatial inhomogeneities
in r
an
disappear) prevail. At the transition
ix x ∼
g
i
4 The time needed to reach locally complete (or almost complete) reactive
equilibrium.
192 MODELS IN TIME AND SPACE
a b
Fad = C v
C Fdiff
Fig. 8.3: (a) The transport in a flow field (Fad , advective transport) is directed,
i.e. it has the same direction for all substances contained in a given volume of
fluid, and an absolute value that is proportional to the various substance con-
centrations C. (b) Diffusive transport results from a large number of random
individual movements. The net flux, Fdiff , is directed from the higher to the lower
concentration (Fick’s first law)
The advective mass flow Fad (transported mass per time and per area
perpendicular to the flow) is:
5 Because C depends on several coordinates (t, x, . . .), we use the symbol for
the partial derivative (∂) in the following discussion, to remind you that when
calculating the derivative the other coordinates are held constant.
194 MODELS IN TIME AND SPACE
Analogous equations are valid for y and z directions. If the cause of the
transport process is thermal movement of the molecules, Dx is called the
molecular diffusion coefficient. Usually, it is the same for all three dimen-
sions (Dx = Dy = Dz ≡ D). Thus the three components of Eq. (8.5) can
be summarized in one vector equation:
Fdiff = −D ∇C (8.6)
∇ is the gradient operator:
⎛ ⎞
∂C/∂x
∇C = ⎝ ∂C/∂y ⎠ (8.7)
∂C/∂z
Eq. (8.5) is called a flux-gradient model (as is Eq. 8.6, of course). The minus
sign shows that transport is always directed against the gradient, i.e. from
the higher to the lower concentration. Wherever undirected random pro-
cesses are involved, flux-gradient models appear, as for instance in heat
conduction (Fourier’s law).
Determining the mixing time τmix,diff for diffusion is somewhat more
difficult than for advection. We could ask ourselves, for instance, how a
concentration jump (a front) in a fluid behaves under the influence of diffu-
sion, or how quickly a substance initially concentrated in one spot spreads
through space (in one, two, or three dimensions). In all these cases we can
express the diffusive mixing distance with the following equation:
Ldiff = a(Dt)1/2 (8.8)
where the numerical factor depends on whether diffusion occurs along one,
two √
or three spatial dimensions. In the case of one-dimensional diffusion,
a = 2. This yields the famous Einstein-Smoluchowski relation6 :
Ldiff = (2Dt)1/2 (8.9)
Note that in contrast to advection, in which the distance
√ L grows linearly
with time t, diffusion only advances proportionally to t since it is undi-
rected and not “goal-oriented”. Solving Eq. (8.8) for t yields the diffusive
mixing time τmix,diff :
1 L2 L2
τmix,diff = = (8.10)
a2 D 2D
where the expression on the far right corresponds to the Einstein-Smolu-
chowski relation for one dimension.
6 In Chap. 8.4.1 we will derive this relation for a special case (see Eq. 8.54)
ADVECTION, DIFFUSION AND EXCHANGE 195
real, for instance between a liquid and a gas. Transport across this boundary
is called gas exchange. At the interface between the gaseous and liquid
phase, the concentration (expressed as mass per volume) exhibits a sudden
jump, i.e. a mathematical discontinuity, so that the concentration gradient
needed for the application of Eq. (8.5) becomes infinitely large.
We thus need a different approach. In Sect. 3.1 we showed that the
equilibrium between two chemical phases A and B can be described by a
static equilibrium relation of the following form (eq = equilibrium):
CA
KA/B = (8.11)
CB eq
eq
Thus, for a given CB there is a value CA = KA/B CB which is in equilibrium
with CB . If A signifies air (A = a) and B water (B = w), then KA/B is
the non-dimensional Henry coefficient Ka/w (Eq. 3.2).
The mass flow across a real boundary (such as the water surface) is
described by the following expression:
eq
FA/B = vA/B (CA − CA ) [M L−2 T −1 ] (8.12)
Room A
FB→A = kB NB − kA NA (8.13)
kB
NAeq = NB (8.14)
kA
This relation has the same form as the exchange flow in Eq. (8.12). For
NA ≤ NAcrit to remain true, NAeq < NAcrit must hold, therefore:
kB
NAeq = NB ≤ NAcrit (8.16)
kA
This condition is fulfilled as long as the following is true:
kA crit 0.1
NB ≤ N = 30 = 10 (8.17)
kB A 0.3
If NB suddenly exceeds this value, the museum staff has time of about
τ ∼ 1/kA = 10 min to bar entry to room A.
It’s important to note that the formalism developed in Eqs. (8.13)–(8.17)
implies a precision and controllability which in reality does not exist. The
visitor behavior which we used to determine the model parameters kA and
kB is only meaningful in a statistical sense. The real values fluctuate around
their mean. We could describe this fluctuation by the standard deviations
σA and σB . When dealing with statistical fluctuations, absolute require-
ments such as “NA must not exceed NAcrit ” often cannot be fulfilled except
through absolute conditions (such as simply keeping room A or B locked).
For real (statistical) systems it is often more appropriate to work with
probabilistic requirements such as “NA shall not exceed NAcrit with a prob-
ability larger than p (e.g. p = 5 % or 0.05)”. The situation then becomes
a typical problem of probability theory. And finally, in order to ease the
demand on museum staff to manage visitor flows, we could of course also in-
stall light barriers at the treasure chamber’s entrance and exit. This would
allow us to constantly monitor NA and trigger a signal as soon as NAcrit is
exceeded.
Boundary layers
Typical case of an
exchange across the
boundary...
Water
Cw
C (z)
processes. With the example of gas exchange we will now show that the
two models are not just physically but also mathematically related.
The so-called film model for the exchange of a volatile substance (sub-
stance with a large Henry coefficient) assumes that on the water side of
the water surface, a thin film exists (the molecular boundary layer with
thickness δ) through which dissolved substances can only travel by molec-
ular diffusion (Fig. 8.5). In the air and in the water beneath the boundary
layer, mixing is so rapid that we can assume the concentrations Ca and
Cw as spatially constant. At the very boundary between air and water, the
eq
two phases are at equilibrium (Cw = Ca /Ka/w , Ka/w = non-dimensional
Henry coefficient). Inside the boundary layer we observe a linear concen-
eq
tration profile from Cw to Cw . The mass flow, according to Fick’s first law,
is then as follows (see Eq. 8.5):
∂C C eq − Cw
Fw→a = −D = −D w
∂x δ (8.18)
D
= vw/a (Cw − Cw
eq
), vw/a =
δ
Thus, the exchange velocity vw/a can also be interpreted by the Fickian
diffusion model. A similar case will be discussed in Example 8.2. More about
modeling the exchange at boundary layers can be found in Schwarzenbach
et al. (2003).
Molecular
In water 10−6 –10−5
In air 10−1
Turbulent, in the ocean
Vertical, surface layerb 10−1 –104
Vertical, deep layer 1–10
Horizontal 10 –108
2
Turbulent, in lakes
Vertical, epilimnionb 10−1 –104
Vertical, hypolimnion 10−3 –10−1
Horizontalc 101 –107
Turbulent, in the atmosphere
Verticald 104 –105
In running water
Turbulent, vertical 1–10
Turbulent, lateral 10 –103
1
a
1 cm2 s−1 = 8.64 m2 d−1
b
Maximal values
c
Horizontal diffusion depends on the considered spatial extent
d
In the atmosphere, horizontal transport primarily occurs by advection
(wind)
average flow. It can be shown that the effect of the fluctuation v on the
mass flow results in an additional undirected transport. Therefore it often
makes sense to describe this so-called turbulent mass flow in the same form
as Eq. (8.5):
∂C
Fturb,x = −Kx (8.19)
∂x
The molecular diffusion coefficient is replaced by the much larger turbulent
diffusion coefficient K. Because the effect of turbulence is usually more
pronounced along the horizontal than the vertical axes (see Sect. 8.1), it
makes sense to add a subscript to K to indicate along which axis turbulent
transport is being considered (i.e. Kx , Ky or Kz ).
The molecular diffusion coefficient is a property of the diffusing sub-
stance and the fluid in which diffusion takes place. It’s usually listed in
physical or chemical manuals. In contrast, the turbulent diffusion coeffi-
cient depends on the medium (lake, ocean, atmosphere, etc.) and on time.
For instance, it makes a significant difference whether we look at turbulent
diffusion in the atmosphere during a calm or windy period, a thunderstorm
or a tornado. Because of the arbitrary distinction between average flow and
turbulence resulting from the averaging interval, the horizontal diffusion in
ocean and atmosphere depends on the spatial extent of the diffusion pro-
cess considered. The figures given in Table 8.1 are typical values; the actual
quantitative determination of turbulent diffusion coefficients can be a very
time-consuming procedure. More about this can be found in Schwarzenbach
et al. (2003), as well as a detailed explanation of the related phenomenon of
dispersion, which appears along the direction of advective flow (for instance
in flowing waters).
10 m
Depth z B
25 m
Lake bed
Kz
Fthermocline = (CA − CB ) = vex (CA − CB )
δ
With:
Kz = 0.6 m2 d−1
δ = 5m
vex = Kz /δS = 0.12 m d−1
CA = 12 mg/L = 12 g m−3
CB = 6 g m−3
(CA , CB : O2 concentration at A and B)
Thus: Fthermocline = 0.12 m d−1 × 6 g m−3 = 0.72 g m−2 d−1
Only after all oxygen in the deep water has completely disappeared will
Fthermocline be of similar magnitude as the loss rate R. In Problem 8.2 we
will deal with the question whether CA does indeed remain constant.
202 MODELS IN TIME AND SPACE
Fx (x) Δx
y
x
If we now take the limit Δx → 0, the right-hand side will be the negative
derivative of Fx with respect to x:
∂C ∂Fx
=− (8.21)
∂t ∂x
In the same way we can calculate the effect on C of the other flow compo-
nents (Fy , Fz ). This ultimately gives us the following result:
∂C ∂Fx ∂Fy ∂Fz
=− + + = −∇F (8.22)
∂t ∂x ∂y ∂z
On the right-hand side is the so-called divergence of the transport
vector F .8 Equation (8.22) is the divergence theorem. It states that the
change of a scalar quantity (concentration C) inside an infinitely small test
volume is equal to the divergence of the vector F describing the flux of
this quantity. In other words: if we know all the fluxes at the surface of
a volume, we know what happens inside that volume (of course assuming
that the volume contains no sources or sinks).
It is remarkable that, to derive the divergence theorem in Eq. (8.22),
we didn’t need any information about how the flux came to be and how
it is mathematically formulated. The theorem holds for any conservative
vector field. We now want to apply it to the advective and diffusive flow
equation9 :
For advection (Eq. 8.3):
∂C ∂ ∂ ∂
= − (Cvx ) − (Cvy ) − (Cvz )
∂t ad ∂x ∂y ∂z
(8.23)
∂vx ∂vy ∂vz ∂C ∂C ∂C
= −C + + − vx − vy − vz
∂x ∂y ∂z ∂x ∂y ∂z
It can be shown that for a non-compressible fluid, the divergence of the
flow field is zero, so the expression in parentheses in Eq. (8.23) disappears.
What remains is:
∂C ∂C ∂C ∂C
= −vx − vy − vz (8.24)
∂t ad ∂x ∂y ∂z
For diffusion (Eq. 8.6):
∂C ∂ ∂C ∂ ∂C ∂ ∂C
=− −Dx − −Dy − −Dz
∂t diff ∂x ∂x ∂y ∂y ∂z ∂z
∂2C ∂2C ∂2C
= Dx + D y + D z
∂x2 ∂y 2 ∂z 2
∂Dx ∂C ∂Dy ∂C ∂Dz ∂C
+ + +
∂x ∂x ∂y ∂y ∂z ∂z
(8.25)
8 Note the different meaning of the symbol ∇ in Eqs. (8.22) and (8.6). In the
latter case, ∇ creates a vector (the gradient vector) from a scalar (C), in the
former case, a scalar (the divergence) from a vector(F ).
9 To do so, we need the differentiation rule for a product of two functions.
204 MODELS IN TIME AND SPACE
If the diffusion coefficients are spatially constant, the last three terms of
Fick’s second law describes Eq. (8.25) disappear. The first three terms then form Fick’s second law of
the diffusion-induced diffusion. Let’s look at the case of isotropic diffusion (Dx = Dy = Dz ≡ D,
change of concentration i.e. diffusion that is equal in all three spatial directions, which is usually
over time. the case for molecular but not turbulent diffusion). In this case, Fick’s
second law takes on the following form, in which it usually appears in the
literature: 2
∂C ∂ C ∂2C ∂2C
=D + + (8.26)
∂t ∂x2 ∂y 2 ∂z 2
Table 8.2 gives an overview of the effects of diffusion and advection on the
flow and local change of a scalar (such as concentration or temperature).
Table 8.2: The effect of diffusion and advection on the flow or the local change of
a property. The equations are formulated for the case of a concentration, C. Note
that when moving through the table each step to the right or downwards results
in a change of sign and an increase of the derivative by one order
2
Diffusion −Dx ∂C
∂x +Dx ∂∂xC2
a
For vx , Dx = constant. Analogous equations exist for the other spatial
coordinates y and z
10 It
makes sense to write the zero-order reaction as a positive and the first-
order reaction as a negative function.
STEADY-STATE MODELS 205
d2 C dC
Dx 2
− vx − kr C + J = 0 (8.29)
dx dx
J
C(x) = A1 eλ1 x + A2 eλ2 x + (8.30)
kr
Dx λ2i − vx λi − kr = 0 (8.31)
1
λi = vx ± (vx2 + 4Dx kr )1/2
2Dx
# 1/2 $ (8.32)
vx 4Dx kr
= 1± 1+ [ L−1 ]
2Dx vx2
The eigenvalues λi determine the spatial structure of the solution, for in-
stance the behavior of C(x) for x → ± ∞. Their dimension is [ L−1 ]. Note
that the eigenvalues don’t depend on the inhomogeneous term J.
The coefficients Ai depend on the boundary conditions of the system.
A second-order equation needs two boundary conditions for its solution to
be unique. If we are looking for the solution of Eq. (8.29) for a finite interval,
e.g. on the x-axis between x = 0 and x = xL , we have six possibilities to
206 MODELS IN TIME AND SPACE
specify the boundary conditions. We can either directly specify the two
boundary values:
C0 ≡ C(x = 0) ; CL ≡ C(x = xL ) (8.33)
Or we can specify the first derivatives at the boundaries:
dC dC
C0 ≡ ; CL ≡ (8.34)
dx x=0 dx x=xL
Finally, we can take any combination of value and derivative. If the system
is unbounded (xL → ∞), the coefficient Ai in Eq. (8.30) which is associated
with an eigenvalue λi with a positive real part must be zero. Otherwise,
the solution C(x) becomes infinite for x → ∞. Explicit expressions for Ai
for all these possibilities are compiled in Table 8.3.
Boundary conditionsa A1 A2
(CL − kJr )−(C0 − kJr )eλ2 xL −(CL − kJr )+(C0 − kJr )eλ1 xL
(1) C0 , CL eλ1 xL −eλ2 xL eλ1 xL −eλ2 xL
−C0 eλ2 xL −CL
+C0 eλ1 xL
(2) C0 , CL
CL
λ1 (eλ1 xL −eλ2 xL ) λ2 (eλ1 xL −eλ2 xL )
CL −(C0 − kJr )λ2 eλ2 xL
−CL +(C0 − kJr )λ1 eλ1 xL
(4) C0 , CL λ1 eλ1 xL −λ2 eλ2 xL λ1 eλ1 xL −λ2 eλ2 xL
(5) C0 0 C0 − J/kr
a
See definitions in Eqs. (8.33) and (8.34). Note that the two missing com-
binations (CL , CL and C0 , CL ) follow from the other four simply by ex-
changing the boundaries (x = 0 becomes x = L and vice versa)
STEADY-STATE MODELS 207
The 222 Rn activity above the sediment falls exponentially towards the
constant value C∞ = JRa /kRn . In Problem 8.4 we will look into how 222 Rn
measurements can be used to determine the vertical turbulent diffusion
coefficient in the ocean.
The λi are the non-dimensional eigenvalues. The solution of Eq. (8.30) is
now:
J
C(ξ = x/xL ) = A1 eλ1 ξ + A2 eλ2 ξ + (8.39)
kr
Since ξ only varies between 0 and 1, the size of λi determines the influence
of the spatial functions exp(λi ξ) on the shape of C(ξ). For instance, if
|λi | 1, then exp(λi ξ) barely varies with ξ and is about 1 everywhere. If,
on the other hand, (−λi ) 1, then exp(λi ξ) drops off from 1 (at ξ = 0)
to 0 over a very short distance. In other words, the spatial variation of this
term is limited to a small zone close to the boundary while over the rest of
the interval the term is zero.
To systematically discuss the different patterns of spatial behavior of
Eqs. (8.30) and (8.39), we rewrite the non-dimensional eigenvalues in terms
of two non-dimensional numbers, Pe and Da:
Pe
λi = [sgn(vx ) ± (1 + 4 Da)1/2 ], i = 1, 2 (8.40)
2
|vx |xL
Pe ≡ ≥0 Péclet number (8.41)
Dx
12 Note that the following considerations are only valid if the extent of the sys-
tem is finite, i.e. if it is enclosed by finite boundaries.
STEADY-STATE MODELS 209
Dx kr
Da ≡ ≥0 Damköhler number13 (8.42)
vx2
In Problem 8.5 we will show that the Péclet number Pe can be seen as a The Péclet and Damköhler
measure for the relative velocity of advection versus diffusion in the interval numbers measure the rel-
xL (for Pe 1, advection is faster than diffusion and vice versa). The ative influence of directed
Damköhler number Da measures the ratio between diffusive and advective transport, diffusive trans-
transport during the average life span of the reactive substance, kr−1 . port, and transformation.
Figure 8.8 gives an overview of the influence of Pe and Da on λi and
schematically shows the shape of the steady-state concentration profile
C(x) between x = 0 and x = xL . We can divide the extreme cases into four
groups (explained in more detail on the next page, following Fig. 8.8).
CL
slow A: purely diffusive
2 reaction
k r <<D/L (with little reaction)
C0
diffusion
0 xL
& reaction
CL
vx2 << Dkr otherwise
fast
(Da>>1) B: reactive-diffusive
reaction
C0
transport 0 xL
& reaction
CL
slow C: purely diffusive
vx2 >>
Dkr (with little advection)
|vx |<< D/L advection
(Da<<1) C0
(Pe<<1)
diffusion 0 xL
& advection
CL
otherwise vx<0 D: advective-diffusive
fast
advection vx >0 (diffusion at boundaries)
C0
0 xL
Sediment
Radium-226
We normalize the concentrations with the lower boundary concentration
CBW . In other words, we set C0 = C(ξ = 0) = 1, CL = 0. We then
212 MODELS IN TIME AND SPACE
Standardized depth ξ
tritium (τ1/2 =12 year) 226
Radium
Absolute depth
is barely influenced by
the upwelling. For the
long-lived radium-226
(τ1/2 =1,600 year)
however, there is a clear
difference between the
profiles with and without 226
Radium
upwelling without upwelling
ξ =0 5km
0 1
Bottom water
vx xL 2 m year−1 × 4 × 103 m
Pe = = = 2.5
Kz 3.2 × 103 m2 year−1
Kz kRa 3.2 × 103 m2 year−1 × 4.3 × 10−4 year−1
Da = = = 0.344
vz2 (2 m year−1 )2
Eigenvalues : λ1 = 3.18, λ2 = −0.68
A1 = −0.0217, A2 = 1.0217
The profile is shown in Fig. 8.10. To illustrate the effect of upwelling, the
corresponding profile without upwelling (vz = 0) is included in the figure.
It becomes clear that, due to upwelling, at a given depth the concentration
is shifted to higher values.
Tritium (3 H)
Boundary conditions: C0 = 0, CL = 1
The most important fluid systems on Earth are the ocean and the atmosphere.
214 MODELS IN TIME AND SPACE
The solutions of Eq. (8.45) depend on the initial and boundary conditions.
We assume that the system is infinite in both directions along the x-axis,
i.e. the substance described by the concentration C can spread along the
x-axis from −∞ to +∞. We’ll now look at two different initial conditions.
We now claim that the solution of Eq. (8.45) with the initial condition
Eq. (8.47) has the following form:
m x2
C(x, t) = exp − (8.48)
2(πDx t)1/2 4Dx t
You can verify that this is indeed a solution of Fick’s second law by inserting
Eq. (8.48) into Eq. (8.45) (see Problem 8.9). C(x, t) is a normal distribution
with the variance:
σx = (2Dx t)1/2 (8.49)
16 Note that for the one-dimensional case, m has the dimension [M L−2 ] (mass
per area perpendicular to the x-axis).
TIME-DEPENDENT SOLUTIONS 215
We assume that the atrazine enters the canal at x = 0 and neglect the
time needed to distribute it homogeneously across the canal’s cross-section.
Its further spread can be described by Eq. (8.45) and the initial condition
Eq. (8.47), where the initial mass per cross-sectional area is:
M 103 g
m= = = 2.5 g m−2
wh 50m × 8m
m2 (2.5 g m−2 )2
t0 = =
2
4πKx Ctol 4π × 0.1 m2 s−1 (10−4 g m−3 )2
x =0 x1/2(t 3)
erfc is the complementary error function (see Appendix E.2), which has
the value 1 for x = 0 and 0 for x → ∞.17 As shown in Fig. 8.12, the solu-
tion (Eq. 8.52) consists of concentration profiles which—starting from the
boundary value C0 at x = 0—intrude further and further into the positive
section. One way to quantify the speed of this process is by following the
migration of the distance x1/2 (t) at which the concentration is half the
value at the edge, i.e., C0 /2. We can calculate x1/2 (t) from the following
equation:
x1/2 (t)
erfc( ) = 0.5 (8.53)
2(Dx t)1/2
The erf and erfc functions are given in tables in most mathematical hand-
books. In such a table we can find that erfc(y) has the value 0.5 if y = 0.48.
Thus:
x1/2 (t)
0.48 =
2(Dx t)1/2
17 Note that in the definition of erfc the spatial variable x appears as the lower
2-dimensional:
2
m(2) x + y2
C(x, y, t) = exp − (8.55)
4(πDt) 4Dt
3-dimensional:
2
m(3) x + y2 + z 2
C(x, y, z, t) = exp − (8.56)
8(πDt)3/2 4Dt
Note that m(2) has dimension [ ML−1 ], while m(3) has dimension [ M ]. The
concentration distributions are circularly or spherically symmetric.
In Example 8.5 we’ve seen that in the one-dimensional case the max-
imum concentration of a point source drops with t−1/2 , i.e. rather slowly.
As we can see from Eqs. (8.55) and (8.56), the coefficient declines with t−1
in the two-dimensional case and with t−3/2 in the three-dimensional case.
Let’s therefore have a last look at the behavior of a pollutant, but this time
for the case of two-dimensional diffusion.
m(2)
Cmax (t) = = Ctol
4πKt
Solving for t ≡ t0 :
Question 8.3: Define typical times for diffusion, advection and transfor-
mation.
Question 8.4: The ratio between mixing time τmix and transformation
time τr determines whether spatial differentiation is appropriate when mod-
eling a system. What is the decision criteria?
Question 8.8: Explain the meaning of the Péclet and Damköhler numbers.
Question 8.11: How must we modify Eqs. (8.55) and (8.56) in order to
also include the influence of advection?
(a) How long will it take until the average concentration in the pond falls
below the value Ccrit = 0.1 µg/L, if the average gas exchange velocity
of tetrachloroethene at the pond’s surface is vW/L = 0.4 m d−1 ? No
other elimination processes of tetrachloroethene play a role in the
pond. Furthermore, you can assume that tetrachloroethene rapidly
becomes completely mixed in the water body.
222 MODELS IN TIME AND SPACE
As a reminder: The O2 flux through the thermocline into the deep water is
initially 0.72 g m−2 d−1 and then rises up to about 1.4 g m−2 d−1 when CB
drops to zero.
Note: As per Eq. (8.37), the “excess activity” of radon, Cexc ≡ C −JRa /λRn ,
is an exponential function. Therefore ln(Cexc ) should be a linear function
1
of height h with slope −(λRn /Kz ) 2 . With λRn known we can therefore
determine Kz . The flux follows from Eq. (8.36).
(b) The Damköhler number (Eq. 8.42) is the square of the quotient of
diffusive and advective transport distance in time τ = kr−1 .
(a) Calculate and draw the steady-state TCE profile from the ground-
water up to the surface assuming that TCE is not degraded in the
soil.
(b) Sketch the steady-state concentration profile of TCE, assuming that
it degraded while migrating upwards through the soil. Compare the
two profiles (a) and (b).
(c) Calculate the steady-state flux of TCE from the groundwater into the
atmosphere per surface area (m2 ) and day for case (a) (no decompo-
sition). Remember that only 15 % of the cross section is available for
the diffusion of TCE through air.
(d) Using Eq. (8.10), estimate the time τstat needed for a steady-state
TCE profile to develop.
(e) The answer to question (d) will show that the time to reach a diffusive
steady state is several years. We should therefore additionally consider
the influence from vertical air flow, as can result from air pressure
differences. This might significantly alter the flux calculated in (c)
and the time calculated in (d). Investigate this question for the case
of a small air current of 0.1 m d−1 . Calculate the modified values for
F und τstat . Calculating the Péclet number may be helpful for this.
Aerobic zone
Anaerobic zone
1 2 3 4
zb
(a) Identify the profiles shown in Fig. 8.13 by the four chemicals A, B, C,
and D.
(b) Derive the analytical solutions for the curves of chemicals A and D.
Use Kz and kr as vertical diffusivity and first-order reaction rate
constant.
1. Flux between surface and deep water, divided into the contribution
of diffusion and advection;
2. The same, but for the flux between bottom and deep water;
3. Radioactive decay.
Note: To calculate the flux per surface, use the standardized concentration
profiles. Choose the sign so that all fluxes into the deep water are counted
positively in the balance while all fluxes out of the deep water appear as
negative numbers. The radioactive decay is also a negative contribution to
the balance.
Note: Use the method from Appendix E.3. The diffusion of the mass element
C0 dξ now proceeds in both directions. Using qualitative considerations,
first make a sketch of the distribution of C(x) through the boundary surface
at x = 0.
Appendix A
List of Symbols
Only symbols that appear more than once are listed here.
A, Ai Area [ L2 ]
B(Vk ) Jacobian matrix, calculated at the fixed point Vk
C Concentration [ ML−3 ]
C0 Initial concentration
C∞ Steady state concentration
CX Concentration of substance X [ MX L−3 ]
Caq Concentration in water [ ML−3 ]
Cin Concentration in the inflow [ ML−3 ]
Cair Concentration in air [ ML−3 ]
Csed Concentration in sediment [ MM−3 sed ]
D Diffusion coefficient
Da Damköhler number (Eq. 8.42)
Fi , F Flux
Fad Advective flux
Fdiff Diffusive flux
Fi,j Mass flux from box j to box i in a multi-dimensional model
[ MT−1 ]
I Inhomogeneous term in discrete models
Jp Zero-order production process
Jl Zero-order loss process
J Zero-order net process
Jin Substance input [ MT−1 ]
Jout Substance output [ MT−1 ]
KA/B Chemical distribution (or partition) coefficient between A
and B
Kd Distribution coefficient between particles and water
[ L3 M−1
sed ]
Ka/w Non-dimensional air/water distribution coefficient (or
non-dimensional Henry coefficient)
Kx Turbulent diffusion coefficient in direction x
Ldiff Characteristic diffusion distance
M Mass [ M ]
N Quantity, e.g. of individuals or atoms
P Coefficient matrix, P = (pi,j )
Pe Péclet number (Eq. 8.41)
Q Flow rate [ L3 T−1 ]
Qex Exchange rate [ L3 T−1 ]
R, Ri External relation
T Period of a periodically fluctuating input [ T ]
V Volume [ L3 ]
VE Volume of the epilimnion [ L3 ]
VH Volume of the hypolimnion [ L3 ]
V Generic system variable
V Vector of the system variable V = (Vi )
Vi i-th system variable in a multi-dimensional model
V0 Initial value of the system variable V at time t = 0
V∞ Steady state value of the system variable V
Indices
i, j,. . . Subscript: variable or box number
(n), (k),. . . Superscript: iteration step of a discrete variable
◦, ∞ Superscript: initial value or steady-state/fixed-point value
Appendix B
B.1 Dimensions
[K] Temperature
[N] Number (of individuals, atoms, etc.)
Occasionally it can make sense to add an index to the dimensions, such as:
If dealing with indexed dimensions, only quantities with the same index
can be reduced.
B.2 Units
From the dimensions we arrive at units by choosing a specific system of
measurement, for instance the metric system. The metric system leads us
to the SI units, which are the international standard and commonly used
in most scientific fields:
The units derived from the basic dimensions usually have their own names:
Formulary
dV
= J(t) − k(t) · V (C.1)
dt
In the most general case, the coefficients J and k can be time-dependent.
J is the inhomogeneous term. The solution is
t
0 −ϕ(t) −ϕ(t)
V(t) = V e +e eϕ(t ) J(t )dt (C.2)
0
Homogeneous Equation: J = 0
The solution for a constant k is
J
V(t) = V∞ + V0 − V∞ e−k·t , with steady state V∞ = (C.7)
k
t
0 −k·t
V(t) = V e + e−k·(t−t ) J(t )dt (C.8)
0
J0 J0
V(t) = + (V0 − )e−kt +
k k
(C.10)
J1 ω J1 ω −kt
+√ sin(ωt − arctan ) + 2 e
2
ω +k 2 k k + ω2
. dVi
Vi = (C.11)
dt
the following system of equations results:
.
V1 = J1 + p1,1 V1 + p1,2 V2
. (C.12)
V2 = J2 + p2,1 V1 + p2,2 V2
If both eigenvalues are real and negative, the model has the following steady
state:
p1,2 J2 − p2,2 J1
a1,0 = V∞
1 =
p1,1 p2,2 − p1,2 p2,1
(C.15)
p2,1 J1 − p1,1 J2
a2,0 = V∞ =
2
p1,1 p2,2 − p1,2 p2,1
The elements of the matrix Ai,j are calculated as follows:
a1,1 = q · (p1,1 − λ2 )(V01 − a1,0 ) + p1,2 (V02 − a2,0 )
a1,2 = −q · (p1,1 − λ1 )(V01 − a1,0 ) + p1,2 (V02 − a2,0 )
(C.16)
a2,1 = q · p2,1 (V01 − a1,0 ) + (p2,2 − λ2 )(V02 − a2,0 )
a2,2 = −q · p2,1 (V01 − a1,0 ) + (p2,2 − λ1 )(V02 − a2,0 )
with
1 1 1
q= = = √ (C.17)
λ1 − λ2 (p1,1 − p2,2 ) + 4p1,2 p2,1
2 Δ
d2 f df
a +b + cf + d = 0, a, c = 0 (C.18)
dx2 dx
d
can be transformed by using g = f + c into the homogeneous equation
d2 g dg
a +b + cg = 0 (C.19)
dx2 dx
It has the solution
1 1
λ1 = −b + (b2 − 4ac) and λ2 = (−b − b2 − 4ac)
2a 2a
The solution for f (x) is therefore
d
f (x) = A1 eλ1 x + A2 eλ2 x − (C.21)
c
236 APPENDIX C
d2 f
a + cf + d = 0 (C.23)
dx2
The approach to the solution is the same, except that λ1 and λ2 are calcu-
lated as follows:
c c
λ1 = − and λ2 = − − (C.24)
a a
d2 g dg
− Pe − Pe2 Dag = 0 (C.27)
dξ 2 dξ
FORMULARY 237
bj1
cj = (b2j1 + b2j2 )1/2 ) ; ϕj = arctg , j = 1, 2 (C.34)
bj2
238 APPENDIX C
Usually, the system variables yj (t) are real functions, which means that the
coefficients bj1 , bj2 and cj are real and the pairs (aj1 , aj2 ) are conjugate-
complex. All these coefficients are calculated from the initial conditions
y1◦ and y2◦ .
b11 = A ; b12 = B
ω (C.36)
b21 = B ; b22 = −ω A
This means that y1 (t) and y2 (t) are
y1 (t) = A cos ωt + B
ω sin ωt (C.37)
y2 (t) = B cos ωt − ω A sin ωt
The solutions, as per Eq. (C.33), are then
B 2 1/2
Aω
y1 (t) = (A2 + ω2 ) sin(ωt + ϕ1 ), ϕ1 = arctg B
(C.38)
2 2
y2 (t) = (ω A + B ) 2 1/2
sin(ωt + ϕ2 ), ϕ2 = −arctg B
Aω
Finally, we can also write the solutions in the form of Eq. (C.29) as
−iωt
y1 (t) = 1/2 (A − i B
ω )e
iωt
+ 1/2 A + i B
ω e
(C.39)
y2 (t) = 1/2 (B + iωA)eiωt + 1/2(B − iωt)e−iωt
The coefficients are, as mentioned, complex numbers.
Appendix D
Eigenvalues
n
dVi
= Ri + pij Vj , i = 1, · · · , n (D.1)
dt j=1
.
In vector and matrix form, and with Vi = dVi
dt , we can write it as
.
V = R + PV
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
R1 V1 p11 ... p1n
⎜ ⎟ ⎜ ⎟ ⎜ .. ⎟
R = ⎝ ... ⎠ V = ⎝ ... ⎠ P=⎝ . ⎠ (D.2)
Rn Vn pn1 . . . pnn
If we multiply Eq. (D.2) on both sides from the left with U−1 and add the
identity matrix E = UU−1 in between P and V on the right-hand side,
then it follows that
.
U−1 V = U−1 R + U−1 P(UU−1 )V (D.4)
or, in terms of Z,
.
Z = U−1 R + ΛZ with Λ = U−1 PU (D.5)
Tk
Zk (t) = Zk∞ + (Zk0 − Zk∞ )eλk t , Zk∞ = − (D.8)
λk
Through the inverse transformation V = UZ we can calculate the time-
dependent solutions Vi (t) from the solutions in Eq. (D.8).
• The diagonal elements pii are negative. All elements outside of the
diagonal, pij with i = j, are positive or zero.
These properties are responsible for the fact that P has one eigenvalue
λ = 0. The corresponding transformed variable U−1 y is the sum of all
yi , that is, the total mass in the system (which is constant). All other
eigenvalues λi are negative, i.e. there is a steady state.
y = −P−1 x (D.16)
Because the pij are still constrained,
< 0 for i = j
pij
> 0 for i = j
Therefore all eigenvalues λi are negative, so that a steady state exists in
this case, too.
242 APPENDIX D
D.2.3 Eigenfunctions
The eigenfunctions of the matrix P are only determined up to a constant
(positive or negative) factor. For instance, they can be written as
We will look at some mathematical functions that are useful for the explicit
time-dependent solution of the diffusion equation. More can be found in
Crank (1975) and in Schwarzenbach et al. (2003, Chaps. 18 and 22).
x0 − σ x0 x0 + σ
G(x)dx = 1 (E.2)
−∞
1 dσ 2
D= (E.4)
2 dt
erf(x = 1) = 1, erfc(x = 1) = 0
erf(x = 0) = 0, erfc(x = 0) = 1
C0
dx x =0 x
Note: the integration boundaries in (E.7) result because the smallest dis-
tance of an element dξ to the point x is given by x, the largest distance by
∞.
Equation (E.7) can be transformed
√ by the substitution of the variable
ξ with η = 2√ξDt and thus dξ = 2 Dt dη:
∞
2C 0 x
C(x, t) = √ exp (−η 2 )dη = C 0 erfc( √ ) (E.8)
π 2 Dt
√x
2 Dt
The function erfc is the complement of the error function (see Sect. E.2):
x x
erfc( √ ) = 1 − erf( √ ) (E.9)
2 Dt 2 Dt
Bibliography
Moore WS, Berrien III, Brasswell BH (1994) The lifetime of excess atmo-
spheric carbon dioxide. Glob. Biogeochem. Cycl. 8:23–38
Odum EP (1971) Fundamentals of Ecology. Saunders, Philadelphia
Pearl R, Reed LJ (1920) On the rate of growth of the population of the
United States since 1790 and its mathematical representation. Proc. Nat.
Acad. Csi. U.S. 6:275–288
Matrix, 82 Prediction, 8, 10
singular, 91 Probability, relations, 28
Mixing, 189
Mixing time, 191, 194 Radioactive decay, 93, 121
Model, 3, 7, 43 Radioactivity, 58
autonomous, 47 Rate
continuous in space, 27 first-order, 44
continuous in time, 26 specific, 44
deterministic, 27 total specific, 52
discrete time, 26, 165 zero-order, 44
dynamic, 23 Reaction
equation, 17 first-order, 91
formation, 10 homogeneous, 92
linear, 19 Reaction rate, first-order, 51
mathematical, 3, 7, 8, 15 Reactor, 50, 82, 85, 90
parameter, 16 Reductionism, 3
scientific theory, 9 Renewal time, 52
static, 16, 33, 50 Retention time, 52
stochastic, 27
Saddle point, 147
Singular matrix, 90
Non-autonomous, 59
Specific rate, 44
Nonlinear model
Stable star, 147
autonomous, 126
Steady state, 49
non-autonomous, 132
Steady-state concentration, 50
with one variable, 126
Strange attractor, 158
with several variables, 143 Structural instability, 109, 151
with three variables, 157 Superposition, linear, 54
Normal distribution, 214, 243 Synergism, 141, 142
System, 4
One-box model, linear, 44 boundary, 5, 10, 15
Oscillation stochastic, 29
creeping, 110 System equation, 19
damped, 108 of a dynamic model, 24
undamped, 88 Systems analysis, task of, 10
System variable, 10, 15
Péclet number, 208
Partial derivative, 144, 193 Thermocline, 99
Phase diagram, 88, 107, 110 Time dependence
Phase doubling, 179 explicit, 46, 59
Phase space, 88, 145 implicit, 46
Population model, 48 Time-discrete model, 165
Predator-prey model, 9, 148 Trace of a matrix, 83, 241
Holling-Tanner, 154, 161 Trajectory, 88, 145
Lotka-Volterra, 148, 151, Transformation, 189, 209
161–163 Transformation time, 191
velocity functions, 150 Transport
with self-interaction, 151 directed, 192
252 INDEX