Introduction To Mathematical Modeling
Introduction To Mathematical Modeling
MATHEMATICAL MODELING
INTRODUCTION TO
MATHEMATICAL MODELING
MAYER HUMI
WORCESTER POLYTECHNIC INSTITUTE
USA
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
© 2014 by Taylor & Francis Group, LLC
CRC Press is an imprint of Taylor & Francis Group, an Informa business
No claim to original U.S. Government works
Printed on acid-free paper
Version Date: 20161101
International Standard Book Number-13: 978-1-4987-2800-3 (Hardback)
This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have been made to publish
reliable data and information, but the author and publisher cannot assume responsibility for the validity of all materials or the
consequences of their use. The authors and publishers have attempted to trace the copyright holders of all material reproduced in this
publication and apologize to copyright holders if permission to publish in this form has not been obtained. If any copyright material has not
been acknowledged please write and let us know so we may rectify in any future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or utilized in any form by
any electronic, mechanical, or other means, now known or hereafter invented, including photocopying, microfilming, and recording, or in
any information storage or retrieval system, without written permission from the publishers.
For permission to photocopy or use material electronically from this work, please access www.copyright.com
(http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923, 978-
750-8400. CCC is a not-for-profit organization that provides licenses and registration for a variety of users. For organizations that have
been granted a photocopy license by the CCC, a separate system of payment has been arranged.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for identification and
explanation without intent to infringe.
Visit the Taylor & Francis Web site at
http://www.taylorandfrancis.com
and the CRC Press Web site at
http://www.crcpress.com
Contents
CONTENTS
1.1 What is Model Building?
1.2 Modeling Framework
1.3 Genes and Biological Reproduction
1.1 What is Model Building?
Definition: modeling is the art of describing in symbolic language a real life system so that
approximately correct predictions can be made regarding the behavior or evolution of the system
under varied circumstances of interest.
We now elaborate on this definition.
First, note that in this definition “modeling” is referred to as an art. As such one cannot develop
rigid preset rules for this task. What can be done, however, is to point out a pattern that is found to be
useful in many cases and can help the practitioner to avoid many pitfalls.
Furthermore, a model is described as being able to make “correct predictions” about the system.
This usually does not mean 100% accuracy. Predictions of many models have a rather wide error
margin. The pertinent question, therefore, is whether these margins are acceptable to the user or not.
Moreover, it might turn out that several models are capable of describing the same phenomena with
different degrees of accuracy (and complexity).
Another important aspect of the definition is that a model should be “solvable.” A sophisticated but
“insolvable” model might be less useful from a practical point of view than a simple and
straightforward one which is capable of making predictions with acceptable error margins.
One should also bear in mind that every model is constructed with certain limitations on its
validity, and these should be borne in mind by the prospective user. Thus, in many practical
applications it is not that “the model is incorrect” but it is the application which Figure 1.1 violates
the basic assumptions of the model used.
Figure 1.1 A flow chart of the modeling process.
A classical example to illustrate these points is given by gravitation theory. Here we do have at
present two concurrent theories which pertain to modeling the same phenomena viz. Newton’s Law of
Gravitation and Einstein’s theory. Though it is accepted and proven that Einstein’s theory is better and
more accurate, still it is highly complex and “hard to solve.” As a result, in most terrestrial
applications we use Newton’s Law of Gravitation with “acceptable error margin.”
As to the problems which require model construction, their source and scope vary between applied
problems in life to attempts to duplicate natural phenomena and (what might seem to be) intellectual
curiosity.
Examples and Illustrations:
1. In many cases of daily life, we construct “mini-models” without even paying any attention to
these facts; e.g., “How do I get to downtown?” (by car, by bus, by subway, on foot, or
otherwise) requires a model which depends on:
The distance to downtown,
The time element (How fast do I want to get there?),
Money considerations,
Security considerations (Is it safe to ride the subway?),
Availability of means (How frequently do the buses run?),
The mood of the person.
2. Consider a truck company operating in the U.S. with “truck depots and service centers” in
some major cities.
A major problem for such a company is how to dispatch trucks to their destinations in the
most economical way (saving gas and drivers’ time).
3. How can the wheat crop be increased to feed the growing human population?
4. What is the cause of global warming and climate change and how can these effects be
mitigated?
5. How can sound and light be recorded in a better way?
6. How can rockets be sent to the Moon or the planets?
7. Why is the sky blue?
where N is the total number of trucks operated by the company (symbolic representation and
constraint).”
Example 4: “For transportation purposes, it is enough to represent the U.S.
map by a set of discrete points whose location coincide with the major cities (abstraction and
simplification).”
STEP 7: Derive the equations that govern the phenomena.
Example 1: If F is acting on point particle of mass m, and the particle acceleration is denoted by a,
then
F=ma.
This is Newton’s Second Law.
Example 2: Denoting by P, V, and T the pressure, volume, and temperature of a gas, then
PV=RT
where R is a constant. This is the Ideal Gas Law.
Remarks: Broadly, mathematical models are classified as deterministic versus stochastic (viz.
probabilistic). Another possible classification of these models is as continuous (i.e., the variables
used are continuous) and discrete. Each of these classifications has its merit within a given context.
STEP 8: Model testing.
To this end, one must solve the model equations and compare the solution with the actual data
collected in steps 2 and 3. If there is a bad fit,
i.e. non-acceptable deviations, then it will be necessary to redo steps 4, 5, and 6.
In this context, we remark that sometimes new mathematical techniques have been devised to solve
a mathematical model. If the model equations remain intractable, then some approximations to the
model equations must be made, thereby sacrificing accuracy in favor of easier computability.
Example: If the original model equations are highly nonlinear and hard to solve, then one may find
an acceptable linear approximation which might be solved easily.
STEP 9: Model limitations and constraints.
At this point, one must become clearly aware of the limitations that must be imposed on the use of
the model and the permissible range of the variables.
Example 1: One cannot use the equations of classical mechanics to predict the motion of a particle
whose speed is close to the speed of light.
Example 2: The ideal gas law is a good model for some gases but not for others.
STEP 10: Predictions and sensitivity analysis.
Once a model has been tested and found acceptable, then it can be used to make predictions.
Whenever such a prediction is found to be correct, the model is considered to be more reliable (in a
way every such prediction is a further test of the model).
One should bear in mind, however, that sensitivity analysis of many models is required before their
actual use; i.e., one has to find the extent to which the model predictions are sensitive to small
variations of the model parameters. We note that some models are “required” to be highly sensitive
while in others insensitivity to such variations is necessary (e.g., if the data contains inherent errors).
Example 1: Ballistic tables.
To construct an “exact” ballistic table, one has to know the exact atmospheric conditions, amount of
charge, geographic altitude, and state of the cannon to be fired. In field conditions, however, these
variables are known approximately at best. Hence, a good ballistic model must be somewhat
insensitive to small variations in these parameters while giving a reasonable prediction about the
range of the shot.
Example 2: Models for physical resonances.
Here, sensitivity is highly desirable especially when several such “close by” resonances are
involved.
Example 3: Chaotic systems.
When the evolution of a system under consideration displays high sensitivity to the initial
conditions, we say that the system is “chaotic.” Under these circumstances, it is possible to make only
“short time” predictions about the state of the system. This is why weather forecasts are accurate only
for a “few days” (at best).
STEP 11: Extensions and refinements.
If a model is found to be correct in some instances but less accurate in others, then a refinement of
it is needed to take care of these exceptions.
Example: The ideal gas law needs such a refinement when the gas molecules are “large” (e.g.,
diatomic gases). The refined model is given by
(P+αV2)V=RT
where a is a parameter which depends on the gas.
STEP 12: Compounding.
Once a correct and reliable model has been established for some phenomena, then related
problems can be modeled by a process of compounding.
Example: Once the equations of motion for the spring-mass system are found, one can compound
the model to systems of several masses and springs.
Finally, we present here a schematic overview of the modeling process.
EXERCISES
E. Fermi was one of the great theoretical physicists in the twentieth century. Some of the following
“mini-model” questions are attributed to him. We offer these here to sharpen the reader’s skill in the
modeling of “real world problems.”
1. Ignoring oceans and such, how long would it take to walk entirely around the world?
2. How much water per year flows in the Mississippi river?
3. How many dump-truck loads would it take to move Mt. Washington in New Hampshire,
USA?
4. Find the dimension of a box that can contain all of the human race (five billion
approximately).
5. What is the linear velocity of Earth around the Sun?
6. How many drops of water are in the Pacific ocean?
7. How many books are in a bookstore?
8. How many atoms are in a cell?
9. How many cells are in the human body?
10. How many light-bulbs burn out in one minute throughout the world?
11. What is the actual volume of material in a solid cubic meter of metal (remember atoms are
made of nuclei and electrons)?
12. How do you buy the best car for your money?
13. How do you buy the best computer for your money?
Results
1. It is possible to obtain pure lines of beans, i.e., beans which by self fertilization will always
produce descendants of the same type. However, these pure lines are prone to disease and
therefore not very desirable from a commercial point of view.
2. Sometimes, a plant from a line producing green smooth beans will produce by self
fertilization some descendants which are wrinkled, etc. (so that appearances might be
deceptive).
3. If we cross pure lines of G-S beans with G-W, we obtain first generation G-S beans only.
However, in the second generation, these G-S beans will give both G-S and G-W beans by
approximate ratio of 3:1.
Subproblem:
Build a model to explain texture only; i.e., assume all beans are Green.
Qualitative Model:
1. A bean carries entities which we shall call “genes” which determine whether it is smooth or
wrinkled. These will be denoted by S and W.
2. Each bean contains two such genes.
3. A bean is smooth if the combination of genes is SS or SW and wrinkled if WW.
4. In cross fertilization, one gene is accepted (independently) from each parent.
Remark: In such a situation where the combination SW is smooth, we shall say that S is
“dominant” with respect to W.
Mathematical Model
Let R(*, *) denote the reproduction function, i.e., the probability distribution of the descendants for
a given pair of parents
R(p1,p2)=(R1,R2,R3)
where pi and p2 represent the parents and Ri, R 2 , and R 3 are the probabilities of SS, SW, and WW
descendants respectively. If P(*) is the probability that a given bean carries a certain gene, we then
have
R1(p1,p2)=P(S|p1)P(S|p2)
R2(p1,p2)=P(S|p1)P(W|p2)+P(W|p1)P(S|p2)
R3(p1,p2)=P(W|p1)P(W|p2)
(In these equations P(S | Pi) represents the conditional probability that the parent Pi contributes the S
gene to the descendant and so on.)
Model Predictions: In the cross fertilization experiment, we started with two pure lines, i.e. pi =
SS, p2 = WW. As a result, our model predicts for first generation descendants:
R1=0R2=1R3=0,
e. all first generation beans are G-S which corresponds to the experimental results. For the
second generation, we therefore have
p1=p2=SW
and hence,
R1=1212=14,R2=12,R3=14,
i.e. 3 4 of the beans are smooth and 1 4 are wrinkled, i.e., a ratio of 3:1.
EXERCISES
1. Predict the results of cross fertilization between
SW and WW beans.
SW and SS beans.
2. (Compounding) Devise a model for beans which takes color into account.
Hint: Each bean will now have four genes; S, W for texture and C, c for color. Suppose that
beans with (SS,CC) are crossed with (ww,cc) (and C is also dominant with respect to c)
and the first generation descendants reproduce by self fertilization. Predict the results for
color and texture of the crop.
3. The following are well known facts regarding blood types in humans:
There are four (major) blood types denoted by A, B, AB and O.
Each blood cell contains two genes which determine the blood type.
The O-gene is regressive with respect to the A and B genes, i.e., AO and BO bloods
are A and B bloods respectively.
A and B genes are of “equal strength.”
In the process of reproduction, each parent donates one gene to determine the blood
type of the descendant.
Use this data to:
1. Give an explicit representation for the reproduction function of this system.
2. Predict the blood type distribution for the descendants to parents with blood types AO and
BO.
Bibliography
[1] R. Aris - Mathematical Modelling Techniques (Ferron-Pitman).
[2] B. Barnes and G.R. Fulford -Mathematical Modeling with Case Studies, 3rd Edition (CRC Press).
[3] Edward A. Bender - An Introduction to Mathematical Modeling (Dover).
[4] C. Dym - Principles of Mathematical Modeling, 2nd Edition (Academic Press).[5] Lin, C.C, and Segal, L.A.,1974, Mathematics
Applied to Deterministic Problems in the Natural Sciences, Macmillan, NY.
[6] Lindsay, R.B., and, Margenau, H.,1955, Foundation of Physics, Dover, NY.
[7] Maki, D.P., and Thompson, M., 2006,Mathematical Modeling and Computer Simulations Brooks/Cole, Belmont, CA, USA
[8] Meerschaert, M. M.,2012 - Mathematical Modeling, 4th Edition, Elsevier, Burlington, MA.
[9] Melnik, R., (Editor), 2015 Mathematical and Computational Modeling: With Applications in Natural and Social Sciences,
Engineering, and the Arts Wiley, Hoboken, NJ
[10] Noble, B.,1967, Applications of Undergraduate Mathematics in Engineering, MAA.
[11] Temam, R. and Miranville, R., 2000 Mathematical Modeling in Continuum Mechanics, Cambrige Univ. Press, Cambridge, UK.
CHAPTER 2
CONTENTS
2.1 The Motion of a Projectile
2.1.1 Approximations and Simplifications
2.1.2 Model
2.1.3 Model Compounding
2.2 Springmass Systems
2.2.1 Data Collection
2.2.2 Approximations and Simplifications
2.2.3 Mathematical Model
2.2.4 Remarks and Refinements
2.3 Electrical Circuits
2.3.1 RLC Circuits
2.3.2 Approximations
2.4 Population Models
2.4.1 Logistic Model
2.4.2 Prototype Model
2.4.3 Data and Approximations
2.4.4 Solution of the logistic equation
2.5 Motion in a Central Force Field
2.5.1 Radial Coordinate System in R 2
2.5.2 Linear Pendulum
2.5.3 Nonlinear Pendulum
2.5.4 A Short Introduction to Elliptic Functions
2.5.5 Motion of a Pr ojectile on a Rotating Earth
2.5.6 A Particle in a Central Force Field
2.5.7 Motion of a Rocket
2.5.8 Multistage Rockets
2.5.9 Control of a Satellite in Orbit
2.6 Greenhouse Effect
2.7 Current Energy Balance of the Earth
2.7.1 Critique of the Model
2.7.2 Humanity and Energy
The behavior and evolution of many scientific and engineering systems are described by equations
which involve unknown functions and their derivatives. These are called differential equations, and
methods for their solution play a central role in many disciplines.
Differential equations are classified as ordinary differential equations (ODEs) and partial
differential equations (PDEs). ODEs are equations which involve only one independent variable
while PDEs involve several independent variables.
To motivate the study of these equations we consider in this chapter problems in various areas
which are modeled naturally by ODEs. For some of these models a solution is possible by elementary
integration methods. For others more elaborate methods are needed.
For all the models presented in this chapter we illustrate the modeling process by adhering as
closely as possible to the modeling framework that was introduced in the previous chapter.
2.1.2 Model
With the approximations delineated above it follows from Newton’s second law that the equation of
motion of the projectile is
(2.1)
md2Xdt2=-mgj
where j is a unit vector in the upward vertical direction. Since the only force acting on the projectile
is in the j-direction, we infer also that its motion is constrained to a plane. Without loss of generality
we can choose this plane to be the x-y plane with x = (x,y), (see Fig. 2.1). (Equation 2.1) is
equivalent then to two scalar equations
(2.2)
x¨=d2Xdt2=0,y¨=d2ydt2=-g
Solution 2.2.1 Let the distance between the center of mass of m 1, m 2 at equilibrium be L (if we
idealize the system and treat m 1, m 2 as point particles, then L is the natural length of the spring).
If these centers of mass at time t are at x 1, x 2 respectively, then either a. x 2 - x 1 - L > 0 or b. x
2 - x 1 - L < 0.
In the first case (a) the spring is stretched beyond its natural length, and hence m 1 is pulled to the
right and m 2 to the left (by Newton’s third law these two forces are equal but in opposite directions.)
Hence, using (Equation 2.23), we have
m1d2x1dt2=k(x2-x1-L)
(2.43)
m2d2x2dt2=-k(x2-x1-L)
Similarly in case (b) m 1 is pushed to the left and m 2 to the right. Since x 2 - x 1 - L < 0, we infer once
again that the equations of motion are given by (2.43). Thus the differential equations which govern
the system are the same in both cases.
We observe that this system is modeled by a system of coupled ordinary differential equations.
Example 2.2.2
Derive the equation of motion for a mass in between two springs which are attached to rigid walls
whose distance from each other is L, as shown in .
Solution 2.2.2 In problems of this type it is natural to use a coordinate system whose origin
coincides with the equilibrium position of the mass (which does not have to be calculated) and
obtain a diffe rential equation for the displacement from this position as a function of time. In fact
for such a displacement x the change in the forces acting on m is given by (using Equation (2.23))
Fig. 2.5.
Figure 2.5 A mass and two springs enclosed by rigid walls
F=-k1x-k2x
(regardless of the sign of x). Hence the desired equation of motion is
(2.44)
mx¨=-(k1+k2)x
Remark 2.2.1
To evaluate the position x eq of m at equilibrium we use the fact that in this state F ext = 0. Hence if
ℓ1, ℓ2 are the natural lengths of the springs and m is treated as a point particle we have
k1(xeq-ℓ1)=k2(L-xeq-ℓ2)
(where we used a coordinate system whose origin is at the left wall of the system). However, note
again that Equation x eq is not needed for the derivation of Equation (2.44).
Example 2.2.3
Derive a model equation for the motion of a mass which is attached to a thin elastic bar and
subject to torsional forces (“twists”).
Solution 2.2.3 To model this problem one must conduct the same type of experiments and make
the same approximations as in the spring‐mass system. For small “twists, “ i.e., when the twist
angle θ is small, the elastic restoring torque due to the bar can be approximated by
(2.45)
T=-kθ
Using Newton’s second law for rotating bodies we then have
(2.46)
Iθ¨+kθ=Text
where T ext is the external torque and I is the moment of inertia of m around the axis of rotation which
is defined as
(2.47)
I=∫Vr2ρ(x)dx
Here r is the distance of x from the axis of rotation and ρ(x) is the density of the mass attached to the
bar.
When frictional forces are also present then for | θ ˙ | ≪ 1 , we have
(2.48)
Ff=-bθ˙
and the equation of motion for m becomes
(2.49)
Iθ¨+bθ˙+kθ=Text
Exercises
1. Find the differential equation which governs the motion of the system shown in Fig. 2.7:
Hint: Apply Newton’s second law to the massless point P at which the two springs are connected.
2. Repeat Ex. 1 for the system shown in Fig. 2.8.
Figure 2.8 A mass connected to two springs in “parallel”
(b) Two masses suspended vertically on springs in the gravitational field of the Earth (Fig. 2.10)
Figure 2.10 Two masses suspended vertically on springs
Hint: Derive the equation of motion in terms of the displacements x 1, x 2 of the two masses from
equilibrium as in example 2.
6. Generalize example 2.2.2 to a system of N masses with N + 1 springs as shown in Fig. 2.11.
2. A loop in an electric circuit is a sequence of circuit elements which start and end at the same
point (i.e. a closed path).
First Kirchhoff law:
The algebraic sum of all the currents at a node is 0. For the node in Fig. 2.13 we have
i1-i2+i3=0.
By convention currents coming to the node are considered positive while those leaving it are
negative.
Second Kirchhoff law:
The algebraic sum of the voltage drops around a loop in an electric circuit is equal to the algebraic
sum of the external voltage sources in the loop.
2.3.2 Approximations
1. We assume that the resistance, capacitance, or inductance of a given electrical component is
independent of the environmental factors (such as temperature, humidity, etc.) and the previous history
of the circuit.
2. Cables connecting circuit components have zero resistance, capacity, and inductance.
3. The passage of an electric current through a cable always involves a leakage which leads to a
loss of electric energy. For short distances one can usually ignore this loss. However, over long
distances (i.e. transmission lines) one must take these losses into account.
With this data and approximations one can in principle analyze any given circuit. We present a few
examples.
Example 2.3.1
RLC Circuit.
A simple RLC circuit is illustrated in Fig. 2.14.
Figure 2.14 RLC circuit
To solve it we first note that there are no nodes in this circuit, and therefore only Kirchhoff’s
second law applies, hence
(2.54)
e(t)=Ri+QC+Ldidt
Differentiating with respect to t and using (2.53) we obtain
(2.55)
dedt=Rdidt+1Ci+Ld2idt2
If d e d t is known, then (Equation 2.55) constitutes a second order (inhomogeneous) differential
equation with constant coefficients for the current i in the circuit.
Example 2.3.2
Wheatstone Bridge.
The circuit shown in Fig. 2.14 is used to measure the resistance R x of a resistor by the use of two
fixed resistances R 1, R 2 and a third, variable one, R 3.
In this circuit e is a low voltage battery and G a galvanometer, i.e. an instrument to measure
currents. Once R x is inserted and the circuit is closed, the operator manipulates R 3 until the current in
G is zero (the circuit is then said to be balanced). Apply Kirchhoff’s second law to the loops ACD
and CBD. In the balanced state we obtain
(2.56)
R1i1-R2i2=0,R3i1-Rxi2=0
hence
(2.57)
R2Rx=R1R3,i.e.Rx=R2R3R1
Example 2.3.3
Multicomponent RLC Circuits
Definition 2.3.1
We say that a resistance R is equivalent to R 1 and R 2 in a given circuit if the replacement of these
resistors by R does not affect the current in the circuit.
Example 2.3.4
From Kirchhoff’s second law it is easy to see that:
1. Two resistors R 1, R 2 in series are equivalent to one resistance R = R 1 + R 2
2. Two resistors R 1, R 2 in parallel (see Fig. 2.16) are equivalent to one resistor with
1R=1R1+1R2.
Resistors in parallel
Example 2.3.5
Find the equivalent resistance of the “infinite” circuit in Fig. 2.17.
Thus we infer
R 0 = 2 R + 1 1 R + 1 R 0 i .e . R 0 = [ 1 + 3 ] R .
Example 2.3.6
DC motor.
A DC motor is an electro‐mechanical system consisting of two circuits the field circuit F, the
armament circuit A -and a shaft (Fig. 2.19.).
Background:
The following information about the electro‐mechanical coupling in this circuit is needed to model
this system.
1. The motion of the shaft causes a potential drop V s in the armament circuit which is proportional
to its angular velocity ω and the field current i F
(2.58)
Vs=c1ωiF
The proportionality constant c 1 in this equation is called the electromechanical constant of the motor.
2. The torque T exerted on the shaft is proportional to i A and i F
(2.59)
T=c2iAiF
Model: A mathematical model for this system can be written down now by applying Kirchhoff’s
second law to the two circuits and Newton’s second law to the shaft whose moment of inertia we
denote by J.
(2.60)
LFdiFdt+RFiF=eF
(2.61)
LAdiAdt+RAiA+c1ωiF=eA
(2.62)
Jdωdt=c2iAiF-c3ω
where C3 represents frictional damping which is proportional to the angular velocity of the shaft.
This is a system of three coupled nonlinear ordinary differential equations of the first order.
Example 2.3.7
Circuits with nonlinear resistors
Some electrical circuits contain “vacuum tubes” (or their solid state equivalents). For these
elements the resistance is a function of the current, e.g.,
(2.63)
R=μ(i2-1)
If we substitute this expression for the resistance in Equation. (2.55), we obtain
(2.64)
Ld2idt2+μ(i2-1)didt+1ci=dedt
Without the forcing term this is equivalent to
(2.65)
x¨+μ(x2-1)x˙+kx=0
(Equation 8.34) is called Van der Pol equation. Although this equation was originally derived to
model electrical circuits in vacuum tubes, it has been used since then to provide a basic model for the
function of nerve cells.
Exercises
1. Prove the statements made in example 2.3.4.
2. Derive model equations for the circuit in Fig. 2.20.
3. What happens if the direct current source in the previous exercise is replaced by an alternating
current source ( t ) = 2 sin ( 2 t ) ?
4. Derive model equations for the circuit in the following figure (Fig 2.21).
Figure 2.21 Generalized RLC circuit
Figure 2.22 Evolution of the normalized fish population in the pool as a function of time for two
initial populations
Example 2.4.1
Predator-prey ecosystem
Consider a lake in which there are two species of fish. The first of these F feeds on plants while the
other P is a predator of F. To write down a mathematical model for this ecosystem we assume that
the fish is consumed by the predator at a rate which is proportional to the population size of the
two species (which we denote also by F, P) . Hence
(2.74)
dFdt=aF-bF2-cFP,b,c>0
As for the predator population we assume that
1. It will become extinct without its prey (rate of death will exceed rate of birth).
2. P increases at a rate which is proportional to F and P (remember as F increases food becomes
more abundant for P).
Thus we infer that
(2.75)
dPdt=-kP+eFP
The system (Equations 2.73)–(2.75) is a special case of Lotka-Volterra equations which model such
ecosystems.
Example 2.4.2
SIR epidemics model
The SIR model for the spread of disease or epidemics is due to W.O. Kermack and A. G.
McKendrick. It assumes that the size of the population remains unchanged (no births and deaths).
Since its inception in 1927 the model has been generalized in various ways. Here we consider only
the original model.
This model assumes that at time t = 0 a part of the population is infected with some infectious
disease. We wish to derive equations for the spread of this disease within the population.
To derive these equations the model divides the individuals within the population into three
groups:
1. S -individuals susceptible to the disease but not infected as of yet.
2. I -infected individuals who are free to mix in the population at large and transmit the disease.
3. R-individuals who contracted the disease but recovered and are no longer susceptible to the
disease.
Subject to the constraint
(2.76)
I+S+R=N
where N is the (fixed) size of the population.
Assuming a free mixing between the S and I groups we infer that:
2.77a
S˙=-aSI
2.77b
I˙=-bI+aSI
2.77c
R˙=cI
where a, b, c > 0.
We now use (Equations 2.77) to demonstrate how in some instances one can derive inferences
about the behavior of a system even without solving the differential equations that govern its
behavior. In particular we show that if the population size is constant then according to this model:
1. b = c.
2. aS(0) < b implies that there will be no epidemics, i.e. I(t) will decrease in time.
3. For all times S(t) > 0, thus the population will always contain some healthy individuals.
To prove the first statement we differentiate (Equation 2.76) with respect to time
(2.78)
I˙+S˙+R˙=0
Substituting Equations (2.77a) - (2.77c) in (Equation 2.78) yields b = c.
To prove the second statement we first observe that S ˙ ( t ) ≤ 0 since S, I ≥ 0 in Equation (2.77a).
Hence S(t) is a decreasing function of time, i.e. S(t) ≤ S(0) . Now from Equation (2.77b) we have
(2.79)
I˙=(-b+aS)I
If aS(0) < b then aS(t) < b for all t. Therefore I ˙ < 0 ( i . e . I ( t ) ≤ I ( 0 ) ) and there will be no
epidemic.
Finally to prove the third statement we use the chain rule
(2.80)
dSdt=dSdRdRdt
Hence (from Equations (2.77a),(2.77b))
(2.81)
dSdR=-abS
Integrating this equation with respect to R we have
(2.82)
S(t)=S(0)e-a/bR>S(0)e-a/bN>0
which proves our statement.
Example 2.4.3
Chemical reactions
The basic law which governs the rate of a chemical reaction is the “law of mass action” which
states that the rate of a reaction is proportional to the (active) concentration of the reactants.
Thus for the reaction
X+Y→ Z
(2.83)
d[Z]dt=k[X][Y]=-d[X]dt=-d[Y]dt
where [X],[Y], [Z] stand for the (active) concentration of the corresponding chemicals.
If the initial active molar concentration of X, Y is a, b, then at time t
[X](t)=a-[Z](t)
(2.84)
[Y](t)=b-[Z](t)
since the production of one mole of Z requires one mole of X and Y. Hence
(2.85)
d[Z]dt=k(a-[Z])(b-[Z])
This differential equation can be solved by direct integration, and if a ≠ b we obtain (using the initial
condition [Z](0) = 0)
(2.86)
k=1t(a-b)lnb(a-[Z])a(b-[Z])
This relationship is usually used to determine experimentally the rate of the reaction. The solution for
concentration of [Z] as a function of time is
[ Z ] ( t ) = b exp [ k ( a - b ) ( t + c 1 ) ] - a exp [ k ( a - b ) ( t + c 1 ) ] - 1 .
If [Z](O) = 0 then
c 1 = ln a b k ( a - b ) .
Example 2.4.4
Catalytic reactions
Of particular interest from a chemical point of view are reactions where the addition of some
“catalyst” accelerates the rate of a reaction that is “slow going.” As an example we consider the
oxidation of sulfur dioxide using nitrogen dioxide as a catalyst.
(2.87)
NO2+SO2→ k1SO3+NONO+12O2→ k2NO2
Observe that the net result of this reaction is
(2.88)
SO2+12O2→ SO3
i.e. the amount of NO 2 in the chemical reactor remains unchanged. Thus a catalyst provides a path
for a desired reaction to happen which has a lower activation energy than the uncatalysed
reaction.
A model for the reactions in (Equation 2.87) consists of five coupled nonlinear differential
equations.
(2.89)
d[NO2]dt=-k1[NO2][SO2]+k2[NO]·[O2]1/2
(2.90)
d[NO]dt=-k1[NO2][SO2]-k2[NO]·[O2]1/2
(2.91)
d[SO3]dt=k[NO2]·[SO2]
(2.92)
d[SO2]dt=-k[NO2]·[SO2]
(2.93)
d[O2]dt=-k2·[NO]·[O2]1/2
Example 2.4.5
Radioactive decay
The nuclei of many isotopes are not stable and therefore decay over time. In many cases the
products of this decay are not stable themselves, and the system then consists of a chain of such
reactions. In all these reactions the decay rate is assumed to be proportional to the “population
size” viz. to the number of nuclei present.
As a particular example we consider here a chain of such reactions where N 1 decays to N 2, which
then decays to a stable nuclei N 3. Here N i , i = 1, 2, 3, represents both the nuclei and their
number.
To model these reactions we consider the time interval [ t , t + △ t ] . On this interval we have
2.94a
N1(t+△ t)N1(t)=-α1N1(t)△ t
2.94b
N2(t+△ t)N2(t)=α1N1(t)△ t-α2N2(t)△ t
2.94c
N3(t+△ t)N3(t)=α2N2(t)△ t,αi>0,i=1,2,3
In Equation (2.94a) the first term on the left hand side represents the number of N 1 nuclei which
were converted to N 2, while the second term represents the number of N 2 nuclei which decayed to
N 3.
Dividing by ▵ t and letting ▵ t → 0 we obtain the system
dN1dt=-α1N1
dN2dt=α1N1-α2N2
(2.95)
dN3dt=α2N2
This is a system of three coupled first order equations. The initial conditions for this system must
specify the number of the nuclei N i , i = 1, 2, 3 at some time t 0.
Exercises
1. Derive a model equation for a fish population which consumes plants as well as itself (Hint:
Remember that a represents the rate of birth minus the rate of death).
2. In a lake there are three species of fish X, Y, Z. X eats plants that are highly abundant. Y is a
predator of X and Z is a predator of X and Y. Derive a model for this ecosystem.
3. Derive a model for an ecosystem which consists of two species X, Y under the following
assumptions:
a. Both species compete for the same nutrient whose supply is limited.
b. There is a migration of X (from outside the ecosystem) at a rate r per unit time.
4. Explain why the coefficient of SI in Equations (2.77a), (2.77b) is the same.
5. Develop model equations for the concentrations of X,Y, and A in the reactions
(2.96)
A+X→ B+2X
(2.97)
X+Y→ B+2Y
(2.98)
A+Y→ B
Observe that the net result of these reactions is 2A → 3B.
6. Show that (Equation 2.86) is the solution of (Equation 2.85).
7. Solve (Equation 2.85) when a = b.
8. Show (by substitution) that the solution of (Equation 2.95) with the initial conditions N 1(0) = N
0, N 2(0) = N 3(0) = 0 is
(2.99)
N1(t)=N0e-α1t
(2.100)
N2(t)=λ2N0[e-α1t-e-α2t]
(2.101)
N3(t)=N0[1-λ2e-α1t+λ1e-α2t]
where λ 1 = α 1 α 2 - α 1 , λ 2 = α 2 α 2 - α 1 .
9. Find the solution of the system, (Equation 2.95), when α 1 ≈ α 2. Assume that N 1(0) = N 0 and N
2(0) = N 3(0) = 0.
10. Let P, Q be the price and quantity of a certain fuel on the open market. A population model for
the evolution of these variables was proposed in the form
P˙=aP/Q-bP2
Q˙=cPQ-dQ2
where a, b, c, d > 0. Justify and discuss the meaning of this model.
2.5 Motion in a Central Force Field
In this section we discuss the equations of motion for a body in a central force field, i.e., when
F = f(r)r where r is the radius vector from the origin and r = |r|. Then various systems (such as the
pendulum) are considered. We begin, however, by introducing the radial coordinate system in R 2.
Using simple trigonometry we infer that the expressions of e r , e θ in Cartesian coordinates are
given by
(2.102)
e r = cos θ i + sin θ j , e θ = - sin θ i + cos θ j
where θ is the angle between the radius vector and the positive x axis.
To obtain expressions for the velocity and acceleration in this coordinate system we observe that
always
(2.103)
x=rer
Hence
(2.104)
X=r˙er+re˙
and
(2.105)
x¨=r¨er+2re˙+re¨
but from (Equation 2.102) we infer that
(2.106)
e˙=θ˙eθ,eθ=-θ˙er
(2.107)
e¨=θ¨eθ-θ˙er
Inserting Equation(2.106)‐(Equation 2.107) in (Equation 2.104)‐(Equation 2.105) we finally obtain
(2.108)
X=r˙er+rθ˙eθ
(2.109)
x¨=(r¨-rθ˙)er+(rθ¨+2r˙θ˙)eθ
Remark: It is useful in some applications to introduce radial coordinates in 3‐dimensions; i.e., attach
to each point in space a triad of orthonormal vectors. If (r, θ, ϕ) are the coordinates of a point in
spherical coordinates (where φ is the angle between the radius vector and z while θ is the azimuthal
angle) then these vectors are
(2.110)
e r = ( sin ϕ cos θ , sin ϕ sin θ , cos ϕ )
e ϕ = ( cos ϕ cos θ , cos ϕ sin θ , - sin ϕ )
e θ = ( - sin θ , cos θ , 0 )
However the expressions for the velocity and acceleration, (Equations 2.108) ‐(2.109) remain
unchanged.
Model: Let x(t) denote the position of the center of mass of m at time t; then in radial coordinates
(2.112)
x=Ler
Since L is constant, we obtain from (Equation 2.109)
(2.113)
x¨=Le¨
x¨=L(θ¨eθ-θ˙er)
Expressing the gravitational force in radial coordinates
(2.114)
F g = - m g j = - m g sin θ e θ + m g cos θ e r
and using Newton’s second law yields
(2.115)
m x ¨ = m L ( θ ¨ e θ - θ ˙ 2 e r ) = - m g sin θ e θ + m g cos θ e r - T e r
where T is the tension in the rod. Rewriting (Equation 2.115) in components form we obtain
(2.116)
L θ ¨ = - g sin θ
(2.117)
m L θ ˙ = T - m g cos θ
The second of these equations can be considered as an equation for T while the first is the desired
equation of motion for the pendulum, i.e.
(2.118)
θ ¨ = - g L sin θ = - g L ( θ - θ 3 3 ! + … )
We observe that (Equation 2.118) is nonlinear in θ. However for small θ we can approximate
(Equation 2.118) by
(2.119)
θ¨+gLθ=0
which is formally the same equation as for the spring mass system without friction. The general
solution of this equation is θ = A cos ( ω t + ϕ ) where A, φ are integration constants and ω 2 = g L .
The period of this pendulum is P = 2 π ω .
Thus if v(t) is the velocity of the rocket and m(t) is its mass we have
(2.149)
[-dmdt△ t](v(t)-u)+m(t+△ t)v(t+△ t)-m(t)v(t)=-m(t)g(h(t))t
where h(t) is the altitude of the rocket at t. Dividing by ▵ t → 0 we obtain
(2.150)
ddt(m(t)v(t))=dmdt(v(t)-u)-m(t)g(h(t))
If we let g(h(t)) be a constant, then the equation of motion reduces to
(2.151)
dvdt=-u1mdmdt-g
Integrating with respect to t we obtain
(2.152)
v = v 0 + u ln m 0 m ( t ) - g t
where m 0 is the initial mass of the rocket and v 0 is its initial velocity. We see that if u ≅ 3 k m / sec
and the flight of the rocket is short, the effect of the last term is negligible and therefore
(2.153)
v = v 0 + u ln m 0 m ( t )
In a one stage rocket
(2.154)
m0=mp+mf+ms
where m p is the mass of the payload, m f the initial mass of the fuel and m s is the mass of the
structures engines and fuel containers. The final velocity of such a rocket (i.e., when the fuel is
exhausted) is given then by
(2.155)
v F = v 0 + u ln m 0 m p + m s
Usually m s m 0 ≈ 1 10 . We see therefore that even if m p = 0 and the effect of
gravity is neglected
v F = v 0 + u ln 10 .
Letting v 0 = 0 and u = 3 we obtain v F ≈ 7.2 km/sec.
Hint: The force exerted by the spring is - k(L - L 0)e r where L 0 is the natural length of the spring.
4. Simulate the system of differential equations that was obtained in the previous exercise with
proper choices for L and k (using “MATLAB” or similar). Compare the period and phase diagrams
(i.e. plot θ ˙ vs. θ) for this pendulum with those for the regular pendulum in Ex 1.
5. Show that when f ( r ) = - μ r 2 in (Equation 2.148) then
1 r = k c 2 ( 1 + e cos ( θ - θ 0 ) )
where e, θ 0 are constants. Discuss the nature of the orbits described by this equation for different
values of e.
Hint: Show that for e < 1 the orbit is an ellipse, e = 1 a parabola, and for e > 1 a hyperbola.
6. Carry out the derivation of (Equation 2.160) from (Equations 2.156)–(2.157)
7. Verify Equations(2.102)‐(2.111).
8. Develop a model for a one stage rocket with air friction. Assume F f = αm(t)v(t) [try to solve
by neglecting gravity]
Hint: d v d t = d v d h d h d t = v d v d h
9. Develop model equations for the motion of the rocket if u = u(t) where u is the speed of the
gases leaving the rocket.
10. Find the optimal value of m 0/m p for a two stage and four stage rocket. What happens when the
number of stages goes to ∞?
Solution
The ratio L = m 0 m p for n‐stage rocket is given by
L=[1-λexp(vFnu)-λ]n
To find the limit of this expression as n → ∞ we rewrite it in the form
exp { ln [ 1 - λ e x p ( v F n u ) - λ ] n } = exp { n ln [ 1 - λ e x p ( v F n u ) - λ ] }
We then have
lim n → ∞ n { ln [ 1 - λ e x p ( v F n u ) - λ ] } = lim n → ∞ ln [ 1 - λ e x p ( v F n u ) - λ ] 1 / n .
As n → ∞ both numerator and denominator in the last expression approach zero; therefore, we can
compute this limit by replacing n (a discrete variable) by x (a continuous variable) and apply
L’Hopital rule:
lim x → ∞ ln [ 1 - λ e x p ( v F x u ) - λ ] 1 / x = lim x → ∞ v F / u 1 - λ e x p ( v F x u ) = v F / u 1
-λ.
Hence
L = exp ( v F / u 1 - λ )
11. A person wants to put a satellite in orbit around the Earth at 64000 km from the Earth’s center. A
short calculation similar to the one in the text shows that the speed of the satellite in orbit should be
(approximately) 2.8 km/sec. Accordingly this person concludes that a one stage rocket is suitable for
this purpose. Explain why this conclusion is wrong.
12. Estimate the effect of gravity on the flight of a one stage rocket if its mass is decreasing linearly
with time:
m = m 0 ( 1 - 5 . 10 - 4 t )
and the flight time is a. 100 sec. b. 190sec. Hint: Use (Equation 2.152) with and without the gravity
term assuming g(h) to be a constant 0 ≤ g ≤ 9.8m/sec 2.
13. Write down model equations for the motion of a rocket which is launched vertically from under
the sea.
Hint: Both below and above sea level the forces acting on the rocket are the thrust T(h), gravity,
and the frictional forces which are proportional to the speed ( o r ( s p e e d ) 2 ) of the rocket. Since
the coefficient of friction is different for h > 0 and h < 0 we obtain two different equations for these
two ranges of h.
14. A person wants to put a satellite in circular orbit around the Earth so that the satellite will
always stay over the same point on Earth. Compute the height and velocity of this satellite (these are
called “geocentric satellites”).
These reservoirs interact dynamically (through diffusion and turbulent mixing) in the same way as
chemical reactants do in a chemical reaction; i.e., we assume that the rate of CO 2 transfer between
the reservoirs is proportional to the concentration of CO 2 in each reservoir. However, not all the
reservoirs connect with each other. If we denote the CO 2 concentration in reservoir i by C i then
(2.164)
dCidt=∑i=17kijCj+Fi,i=1,…,7
where k ij are constants and F i is the forcing (in this model F i = 0 except for the lower atmosphere).
Thus the model depends on 49 parameters (some of which are zero). The proper estimation of these
coefficients (which remains an outstanding research issue) is one of the important steps that is needed
to make accurate predictions about the concentration of CO 2 in the atmosphere.
Once we determine the (mean) concentration of CO 2 in the lower and upper atmosphere, another
model is needed to gauge the impact of this concentration on the mean temperature of Earth (see
chapter 11).
CONTENTS
3.1 Review
3.1.1 Linear differential equations with constant coefficients
3.2 Review of Linear Algebra
3.2.1 Eigenvalues and Eigenvectors
3.3 Reformulation Of Systems Odes
3.4 Linear Systems With Constant Coefficients
3.5 Numerical Solution Of Initial Value Problems
3.5.1 Euler Algorithm
3.6 Finite Difference Approximations
3.6.1 Extension to Higher Dimensions
3.7 Modified Euler And Runge-Kutta Methods
3.7.1 Modified Euler Algorithm
3.7.2 Runge-Kutta Methods
3.8 Boundary Value Problems
Differential equations, which relate a set of unknown functions with their derivatives, play an
important role in many applications of science and engineering. In this chapter, we present some basic
analytical and numerical methods for the solution of some classes of systems of differential equations.
However, our treatment is not comprehensive. We start with a short review of some basic theory.
3.1 Review
A Particular Solution
If the general solution of the homogeneous (Equation 3.2) (with f(x) = 0) is known, then a particular
solution to (Equation 10.1) (with f(x)) ≠ 0 c a n be found by “variation of coefficients.” The
following (short) table summarizes the trial function one has to use to find y p for some important
cases which appear in applications (this table can be extended). The coefficients A, B, and etc. that
appear on the right hand side of the table have to be determined by substituting the trial function in the
differential equation. As a result, one obtains a system of algebraic equations which has to be solved
for the coefficients.
Table 3.1 Finding y p
f(x) yp
ae kx sin β x , cos β x Ae kx A cos β x + B sin β x
a n x n + a n-1 x n-1 + … + a 0 Anxn+ An- 1Xn- 1+ … + A0′
An exception to this table happens when the forcing function f(x) is one of the solutions of the
corresponding homogeneous equation (this is referred to as “resonance”.
Remark
Observe that in general y p will not satisfy the initial conditions. It is only y G that has to satisfy these
conditions.
We illustrate the use of this table by examples.
Example 3.1.3
f(x) = h = constant in Equation ( 3.2 ).
In this case, y p is given by
(3.17)
y p = h c c ≠ 0 h x b + γ 1 b ≠ 0 , c = 0 h x 2 2 a + γ 1 x + γ 2 b = c = 0 ( 3.17 )
where γ 1 and γ 2 are arbitrary constants.
Example 3.1.4
Consider Equation ( 3.2 ) with b = 0, a > 0, c > 0 and
(3.18)
f ( x ) = b 1 cos ω x + b 2 sin ω x .
The general solution of the homogeneous equation is
y h = C 1 cos ν x + C 2 sin ν x
where ν = c a is referred to as the natural frequency of the equation (the frequency of the
oscillations when the forcing function f(x) = 0). If ω ≠ ν, then to find y p we try
(3.19)
y p = A cos ω x + B cos ω x .
Substituting this in Equation ( 3.2 ) we obtain an algebraic equation for A and B
(3.20)
sin ω x [ - a A ω 2 + c A - b 1 ] + cos ω x [ - a B ω 2 + c B - b 2 ] = 0 .
However, since cos ω x and sin ω x are independent functions, we infer that to satisfy this equation
each of the brackets in Equation ( 3.20 ) must be zero. Hence, we obtain the following system of
equations for A and B
(3.21)
-aω2+c00-aω2+cAB=b1b2
since ω is not the natural frequency of the system ( - aω 2 + c) ≠ 0 Equation ( 3.21 ) yields
nontrivial solutions for A and B.
Example 3.1.5
Consider (3.2) with b = 0, a > 0, c > 0 and f(x) = b 1 cos ω x + b 2 sin ω x but with ω = ν.
We refer to this situation where the forcing function has frequency equal to the natural frequency of the
system as “resonance.” (The forcing function f(x) is one of the solutions of the homogeneous equation)
In this case, we use for y p a trial solution of the form
(3.22)
y p = x ( A cos ω x + B sin ω x ) .
Substituting this expression in (Equation 3.2) we obtain
cos ω x [ ( 2 a ω B - b 1 ) + ( - a ω 2 + c ) x A ]
(3.23)
sin ω x [ ( - 2 a ω A - b 2 ) + ( - a ω 2 + c ) x B ] = 0 .
Since cos ω x and sin ω x are independent functions, each bracket in (Equation 3.23) must be zero.
Moreover, since ω is the natural frequency of the system - aω 2 + c = 0,
(3.24)
2aωB-b1=0,2aωA+b2=0.
We obtain the following solution for A and B.
A=-b22aω,B=b12aω
c.A=210-1-2-1012,d.A=111000-1-23
To compute y 1 = y(x 1), we now employ a first order Taylor expansion around x 0 This leads to
(3.56)
y1≅ y(x0)+hy′(x0)=y0+hf(x0,y(x0))=y0+hf(x0,y0).
We now use this (approximate) value of y(x 1) as an initial value for the differential equation on the
interval [ x 1 , x 2 ] . Using the same strategy as before, we obtain
(3.57)
y2=y(x1)+hy′(x1)=y1+hf(x1,y1).
Continuing in this manner, we derive the general formula
(3.58)
yk+1=yk+hf(xk,yk),k=1,…,n-1.
This is known as the Euler algorithm for the numerical solution of differential equations.
Example 3.5.2
Use the Euler algorithm to solve
(3.59)
y′=x+y2,y(0)=0
on the interval [0, 1 ] with step size of 1 3 (see Fig. 2.2). Use linear interpolation to find an
approximate value for y(0.5)
We now show how this algorithm can be extended to systems of equations. To this end, we
consider the system
(3.65)
y′=F(x,y),y(0)=c
where
(3.66)
y=y1⋮ yn,F=f1⋮ fn
and f i = f i ( x , y 1 , … , y n ) on [ a , b ] .
As before, we subdivide the interval [ a , b ] into sub‐intervals of step size h and apply a first
order Taylor expansion of y i on each of the sub‐intervals [ x k , x k + 1 ] :
(3.67)
yi(xk+1)=yi(xk)+hyi(xk)=yi(xk)+hfi(xk,yk1,…,ykm),i=1,…,m.
This can be rewritten transparently using vector notation as
(3.68)
yk+1=y(xk+1)=yk+hF(xk,yk).
This is the vector version of (Equation 3.58).
Example 3.5.4
Solve the following system numerically by using Euler algorithm:
u′=u+2v
v′=u-v,u(0)=0,v(0)=1
on [0, 1] with h = 0.25.
Solution 3.5.4 Applying Equation (3.68), we infer that
uk+1=u(xk+1)=uk+h(uk+2vk)
vk+1=v(xk+1)=vk+h(uk-vk).
Applying this recursively, we obtain
u 1 = u ( 0.25 ) = 0 + 0.25 ( 0 + 2.1 ) = 0.5
v 1 = v ( 0.25 ) = 1 + 0.25 ( 0 - 1 ) = 0.75
u 2 = u ( 0.5 ) = 0.5 + 0.25 ( 0.5 + 2 · 0.75 ) = 1
v 2 = v ( 0.5 ) = 0.75 + 0.25 ( 0.5 - 0.75 ) = 0.6875
u 3 = u ( 0.75 ) = 1 + 0.25 ( 1 + 2 · 0.6875 ) = 1.594
v 3 = v ( 0.75 ) = 0.6875 + 0.25 ( 1 - 0.6875 ) = 0.766
u 4 = u ( 1.0 ) = 1.594 + 0.25 ( 1.594 + 2 · 0.766 ) = 2.376
v 4 = v ( 1.0 ) = 0.766 + 0.25 ( 1.594 - 0.766 ) = 0.973
Exercises
1. Solve the following equations numerically by using the Euler algorithm and compare with
the exact solution. Graph your results.
y ′ = 2 x + y o n [ 0 , 1 ] , y ( 0 ) = 1 , h = 0.2
y ′ = y - y 2 o n [ 0 , 1 ] , y ( O ) = - 1 , h = 0.2
y′ = xy —l on[1, 2], y(1) = 0.5, h = 0.2
y ′ = ( 2 + y ) y - x o n [ 0 , 1 ] , y ( 0 ) = 0 , h = 0.2
y ′ = y sin x on [ 0 , π ] , y(O) = 1, h = π 4 .
2. Solve the equations in Ex. 1 with h 2 and compare with those obtained previously.
3. Solve the equations in Ex. 1 using the Taylor series expansions. Estimate the number of
terms needed to obtain an accuracy of 10−3.
4. Solve
y ′ = sin x + y 2 , y ( 0 ) = 1
on [0, 1 ] with h = 0.25 and use interpolation to find y ( 2 3 ) . Next, solve this equation with
h = 1 3 and compare the two values obtained for y ( 2 3 ) .
5. Use a computer to solve the equation in Ex. 1 with h = 0.01 and h = 0.001 and compare
with the exact solution.
6. Solve the equation in Ex. 4 with h = 0.1 and h = 0.01 to find a better approximation for y ( 2
3).
7. Solve the following systems numerically by using the Euler algorithm
with the initial conditions u(0) = 0, v(0) = - 1 on [0,1] and h = 0.2. Can you solve these
systems analytically?
Hint: Eliminate v from the second equation by using the first equation.
u ’ = u + 2v, v ’ = 4u - 3v
u ’ = u + v, v ’ = u - v + 2x
8. Use the “extended Euler method” where
to solve
y′ = x 2 + 2y, y(0) = 1, on [0,1]
with h = 0.2. Compare with the solution obtained from the Euler algorithm with the same
step size and the exact solution.
yk+1=yk+hyk′+h22yk″
9. Estimate the cumulative error in the “Extended Euler Method” in terms of h.
3.6 Finite Difference Approximations
Before discussing improvements to the Euler algorithm and the solution of boundary value problems,
we must first introduce the notation of O(h k ) (pronounced as “O‐big” of h k ). Moreover, we shall
derive finite difference approximations to the derivatives of a function f(x) in terms of its values at
some set of discrete points.
Definition 3.6.1
Let f(x) and g(x) be two functions which are continuous at x = 0.
We say that f is O(g) if
(3.69)
0 < | lim x → 0 f ( x ) g ( x ) | < ∞ .
Example 3.6.1
sin x = O ( x ) since
(3.70)
lim x → 0 sin x x = 1 .
Example 3.6.2
f = 104 x 3 + 3x 2 + x is O(x 3) since
(3.71)
lim x → 0 f ( x ) x 3 = 10 4
Example 3.6.3
x 4 + x 5 is not O(x 3) since
(3.72)
lim x → 0 x 4 + x 5 x 3 = 0 .
Example 3.6.4
The remainder of the Taylor expansion of f(x + h) around x
(3.73)
Rn(h)=f(x+h)-∑k=0nf(n)(x)hkk!
is O(h) if f (n+1)(x) ≠ 0.
In fact, the remainder of the Taylor expansion is
(3.74)
Rn(h)=∑k=n+1∞f(k)(x)hkk!=hn+1{f(n+1)(x)(n+1)!+hf(n+2)(x)(
n+2)!+…}
(3.75)
lim h → 0 R n ( h ) h n + 1 = f ( n + 1 ) ( x ) ( n + 1 ) ! .
Therefore, we can write
(3.76)
f(x+h)=∑k=0nf(k)(x)hk
From the definition we now infer the following:
Theorem 3.6.1
If f is O(h n ) , g is O(h m ) , and m ≤ n then
1. f ± g is O(h m )
2. f ⋅ g is O(h n+m )
3. f/g is O(h n-m ) .
Example 3.6.5
The local error in the Euler algorithm (i.e., the error in each step) is O(h 2) while the cumulative
(global) error is O(h) .
In fact, the basic approximation in the Euler method is
(3.77)
yk+1=yk+hyk′ +O(h2).
Since there are n steps and n = b - a h , the global error is 1 h O ( h 2 ) = O ( h ) .
The notation of O‐big is used to give the user a rough idea about the behavior of the error committed
by an algorithm as h → 0. This helps to compare between different algorithms and determines which
is superior.
We now introduce finite difference approximations.
If a symbolic representation of a function f(x) is not known, then we consider the function to be
“known numerically” on a domain D if its values on a grid of points in this domain are known and if
by using these values we can calculate an acceptable approximation to the value of u at any other
point in D (by interpolation). It follows, then, that in order to solve a differential equation on D, we
have to introduce a grid of points on this domain and cotruct an algorithm to calculate the values of
the unknown function u at these points. Thus, if (in one dimension) x i , i = 1,...,m, are the grid points,
then the fundamental unknowns that have to be computed are
(3.78)
(fi,Xi)
Since the differential equation and the initial (or boundary) conditions at hand are the only means to
compute these quantities, we must develop an approximation scheme for the derivatives of f in terms
of these data. To see how this can be done, consider the one‐dimensional case where, to begin with,
we assume that the x i ′ s are equispaced (See Fig. 3.3):
xi-xi-1=h
(where h is constant for all values of i).
Thus, suppose that we are given ( x i , f i ) , i = 0, \ldots , n, where x i+1 - x j = h = const. To find an
approximation for f i ′ = f ′ ( x i ) , we use the Taylor expansion of f around x i :
(3.79)
fi+1=f(xi+h)=fi+hfi′+h22fi″ +h33!fi′′′+h44!fi(4)+….
Hence,
(3.80)
fi′=fi+1-fih-h2fi″ -h23!fi′′′-…
or
(3.81)
fi′ =fi+1-fih+O(h).
This formula is called the forward difference approximation for f i ′ (as the point x i+1 is used).
Similarly, we can derive a backward difference approximation for f i ′ using
(3.82)
fi-1=f(xi-h)=fi-hfi′+h22!fi″ -h33!fi′′′+h44!fi(4)….
This leads to
(3.83)
fi′ =fi-fi-1h+O(h).
In both of these approximations, the error is O(h) . However, we can obtain a better approximation by
using “central difference” (i.e., use an interval where x i is at the center). Thus, by subtracting
(Equation 3.82) from (Equation 3.79) we obtain
fi′=fi+1-fi-12h+2h23!fi′′′+25!h5fi(5)+…
or
(3.84)
fi′ =fi+1-fi-12h+O(h2)
We note that although this is a superior approximation, it cannot be used at the ends of the
computational interval ( i . e . , a t x 0 a n d x n ) .
Similar approximation can be derived for the second order derivatives. In fact, adding (Equation
3.79) and (Equation 3.82) leads to
(3.85)
fi″ =fi+1-2fi+fi-1h2+O(h2)
This is a central difference formula. A forward difference formula for f i ″ can be derived by using f
i+2. In fact,
(3.86)
fi+2=f(x1+2h)=fi+2hfi′+2h2fi″ +4h33fi″ +….
Subtracting (Equation 3.79) multiplied by 2 from (Equation 3.86) leads to
(3.87)
fi″ =fi+2-2fi+1+fih2+O(h).
Finite difference approximation for higher order derivatives can be obtained in a similar fashion. For
instance, we have the following central difference formula for the 4th derivative
(3.88)
fi(4)=fi+2-4fi+1+6fi-4fi-1+fi-2h4+O(h2).
Figure 3.3 Equispaced computational grid
Example 3.6.6
Suppose that the values of f at x 0, x 0 + h, and x 0 + 3h are given. Develop a finite difference
approximation for f ’(X 0) with O(h 2) error.
Solution 3.6.1. From the Taylor expansion around x 0 we have
f(x0+h)=f(x0)+hf′(x0)+h22!f″ (x0)+h33!f′′′(x0)+…
f ( x 0 + 3 h ) = f ( x 0 ) 3 h f ′ ( x 0 ) + 9 h 2 2 ! f ″ ( x 0 ) + 27 3 ! f ′ ′ ′ ( x 0 ) + …
To obtain an O(h 2) approximation for f ’(x 0), we must find a combination of f(x 0 + h) and f(x
2
0 + 3h) which cancels the h terms in these expansions. Thus,
9f(x0+h)-f(x0+3h)=8f(x0)+6hf′(x0)-3h3f′′′(x0)+…,
i.e.
f′ (x0)=9f(x0+h)-f(x0+3h)-8f(x0)h+O(h2).
When the grid points are not equispaced, the general technique described above can be easily adapted
to derive appropriate approximations. Thus, if
xi-xi-1=hi,
then
(3.89)
fi+1=f(xi+hi+1)=f(xi)+hi+1f′ (xi)+O(h2)
fi-1=f(xi-hi)=f(xi)-hif′ (xi)+O(h2),
which yields
(3.90)
fi′ =f′ (xi)=fi+1-fi-1hi+hi+1+O(h).
Similarly, for the second‐order derivative,
(3.91)
fi″ =2hifi+1+hi+1fi-1-(hi+hi+1)fi(hihi+1)(hi+hi+1)+O(h).
From (Equations 3.90) and (3.91), we infer that in general the accuracy of our difference
approximations is better with an equispaced grid. (Equations 3.90) and (3.91) or their generalizations
to higher dimensions are, therefore, useful near irregular boundaries where their use is mandatory.
Therefore,
(3.93)
(∂u∂x)ij=ui+1,j-ui-1,j2h+O(h2)
(∂2u∂x2)ij=ui+1,j+ui-1,j-2uijh2+O(h2).
The corresponding formulas for ( ∂ u/ ∂ y) ij and ( ∂ 2 u/ ∂ y 2) ij should be obvious.
Exercises
1. Derive a backward finite difference formula for f i ’’
2. Prove that
f i ′ = - f i + 1 + 8 f i + 1 - 8 f i - 1 + f i - 2 12 h + O ( h 4 ) .
3. Prove the finite difference formula for f i ( 4 ) which is given in (Equation 3.88).
4. Derive an approximation for f ’’(X) with O(h 4) error using f i+2, f i+1, f i , f i-1 and f i-2. (Use a
symbolic computation package if available).
5. Derive a forward difference approximation for f ’’’(X) with O(h) error using f i+3, f i+2, f i+1
and f i
6. Derive a central difference approximation for f ’’’(X) with O(h 2) error.
7. Derive a central difference formula for f ’(x i ) with O(h 2) error if the grid points are not
equispaced, i.e., h i = x i+1 - x i are not constant.
Since the function y(x) in general is not linear, it stands to reason that a better approximation for
y(X) can be obtained if an “averaged slope” of f(x) over [ x i , x i + 1 ] is used. An appealing way to
implement this will be to average the slopes at x i and x i+1, i.e.,
(3.95)
yi+1=yi+h2[yi′+yi+1′]
The only obstacle for the use of this formula is that
(3.96)
yi+1′=f(xi+1yi+1)
This implies that in order to compute y i + 1 ′ we must know y i+1 first!
To circumvent this difficulty we can utilize the Euler algorithm to obtain a predictor for the value
of y i+1
(3.97)
yi+1(p)=yi+hf(xi,yi).
This value can be used then in eq. (3.95) to obtain “corrected” value of y i+1
(3.98)
yi+1(c)=yi+h2[f(xi,yi)+f(xi+1,yi+1(p))]
Should the difference between y (p) and y (c) be larger than some preset error tolerance ɛ, we can use y
(c) as a new y (p) and iterate the process until
(3.99)
|yi+1(p)-yi+1(c)|<ε
(Obviously we must set a limit on the number of these iterations to avoid the possibility of an infinite
loop.)
We illustrate the implementation of this algorithm through the following example.
Example 3.7.1
Solve
(3.100)
y′=2x-y,y(0)=1
with h = 0.2 on [0, 1 ]. Use error tolerance of ɛ = 0.05 and one predictorcorrector iteration.
Solution 3.7.1. Using Equation (3.97), (3.98) we obtain Ta ble 3.4 for the solution of Equation (
3.100 ). To gauge the accuracy of the numerical solution we present in the last column of this table
the values of the exact solution
(3.101)
y=3e-x+(2x-2).
We demonstrate now that the modified Euler algorithm (without iterations) is equivalent to the use of
second order Taylor expansions for y(x) on each of the sub‐intervals [ x i , x i + 1 ] . In fact according
to this scheme
(3.102)
yi+1=yi+hyi′ +h22yi″ +O(h3).
Replacing y i ″ = ( y i ′ ) ′ by its forward finite difference approximation (eq. (3.79)).
(3.103)
yi″ =(yi′ )=yi+1′ -yi′ h+O(h)
leads to
(3.104)
yi+1=yi+h{yi′ +h2[yi+1′ -yi′ h+O(h)]}
(3.105)
+o(h3)=yi+h2{yi′ +yi+1′ }+O(h3).
This also shows that the local error in the modified Euler algorithm is O(h 3) . Consequently, this
leads to the following estimate for the cumulative (global) error
(3.106)
E=nO(h3)=b-ahO(h3)=O(h2).
v′=u+v
on [0, 100 ] using different step sizes. Compare with the exact solution.
Remark 3.7.1
Problems like this where numerical solution diverges from the correct solution are called “stiff”
where we had to add a fifth (fictitious) point y -1 = y( - 0.25) . (Equations 3.116), (3.119) form a
system of four equations in five unknown. To add one more equation we use (Equation 3.83) to
approximate the boundary condition y ˙ ( 0 ) = 1 .
This leads to
(3.120)
y 1 - y - 1 0.5 = 1 .
Together (Equations 3.116), (3.120) form a system of five equations in five unknowns whose solution
is
y 0 = 0.619 , y 1 = 0.412 , y 2 = - 0.255 , y 3 = - 0.123 .
Exercises
1. Solve the boundary value problem given by (Equation 3.113) with h = 10−1, 10−2. What can
be said about the coefficient matrix for the resulting system of equations?
2. Compare the exact and numerical solution of
u″ +2u′+u=xu( 0) =1,u( 1) =3
Stability Theory
CONTENTS
4.1 General Introduction
4.2 Two species model
4.2.1 Steady States
4.2.2 Stability Analysis
4.3 Basic concepts
4.4 Linearizable Dynamical Systems
4.5 Linearizable systems in two dimensions
4.6 Liapounov method
4.7 Periodic SOLUTIONS(LIMIT cycles)
4.1 General Introduction
In the 19th and early 20th centuries most mathematical models for physical phenomena used linear
differential equations which can be solved by analytic techniques. If a model turned out to be
nonlinear, then some methods were used to “linearize” the model to obtain approximate solutions.
With the rapid advancement of science and engineering and the advent of computers, it was found
advantageous to consider sophisticated models where nonlinearities and their impact cannot be
ignored. However while computers can generate “heaps” of numbers, it is not easy to extract insights
about the system evolution under different conditions from this data.
Stability theory addresses the issue of model nonlinearities by making the observation that in many
models an initial state of the system will evolve into a steady state (i.e., a state that is time
independent). It was found that for many practical purposes the transient states of the system (i.e.,
those that are time dependent) are of “marginal” importance while the steady states and their
properties are of utmost practical importance.
Since most mathematical models provide only an approximation to reality, a basic issue about the
steady states is their “response” to “deviations” or perturbations of the system from the steady state.
In other words if the system “somehow” deviates from the the steady state, will it return to it or”run
away from it As an example consider a car running smoothly on a highway then hits a pothole on the
road. This will cause the body of the car to vibrate and the question arises then whether these
oscillations will “damp out” and the car return to its “normal steady state”, or a breakdown will
occur and the car settle (after a short transient state) into a new steady state (i.e., the car is not
drivable and has to be repaired) which is “far” from the original state. In other words “how stable
was the original state of the car to perturbations” The anawer to this fundamental question is the
subject of stability theory.
From another point of view one has to remember that mathematical models contain parameters that
have to be evaluated experimentally and hence, are subject to measurement errors. As a result the
actual steady state of a given system might be somewhat different from the one predicted by the model
equations. The question arises again as to whether these “small deviations” are “benevolent” (i.e.
have little or no practical impact) or “fatal.”
Historically the study of stability theory was initiated by H. Poincare and A. M. Liaponouv.
Poincare was interested in the stability of the solar system and in particular whether the Earth orbit is
stable. That is whether the Earth eventually “fall” into the Sun or recede from it. (This is still an open
question.) A. M. Liaponouv considered the stability of various mechanical systems in particular under
the influence of gravity. Today stability theory is still an important and evolving branch of science and
engineering. In this chapter we give an exposition of the basic ideas of this theory.
To motivate our study and introduce some of the basic ideas we revisit the population model for the
number of fish in a pool (see Chapter 2 Sec. 2.4) and assume that the size N(t) of the population in
this (simplified) ecological model is given by
(4.1)
dNdt=(N-1)(N-3),N(O)=N0
where the quadratic term in N represents the competition for resources (including food).
At a steady state d N d t = 0 and hence Equation (4.1) has two steady (or equilibrium) states N = 1
and N = 3.
To study the stability of these steady states we note that Equation (4.1), though nonlinear, can be
integrated by elementary methods and we obtain
(4.2)
N(t)=(N0-3)e2t-3(N0-1)(N0-3)e2t-(N0-1).
From Equation (4.2) we infer that whenever N 0 ≠ 0 the population N(t) will approach the steady state
N = 1 as t → ∞ even if N 0 is very small or large. (However, N 0 must be positive) . Thus, a (small)
perturbation of this ecological system from N = 1 will cause its state to “come back” to this
equilibrium while similar deviations from N = 3 will cause the the population to “run away” from this
steady state and converge to N = 1 as t → ∞. We conclude, therefore, that the steady state N = 1 is
(asymptotically) stable while N = 3 is unstable.
To demonstrate these results graphically we plotted in Figure 4.1 the solutions to (4.2) with
different initial conditions.
The objective of stability theory (and phase space methods) is to derive these results about the
stability of the steady states without solving analytically or numerically the model equations under
consideration. This is a crucial feature since most current mathematical models lead to systems of
nonlinear differential equations which cannot be solved analytically in closed form. Furthermore,
numerical algorithms for the solution of these equations consume inordinate amounts of CPU time and
in many cases do not converge or converge to the wrong solution.
The underlying idea of these techniques is to consider the model equations as an algebraic
relationship between the state variables and their derivatives. Thus Equation (4.1) is viewed as an
algebraic relation between N and dN/dt. To analyze the stability of the steady states we have only to
remind ourselves (from elementary calculus) that d N d t > 0 implies that N is increasing while d N d
t < 0 implies that N is decreasing with time. Thus in the “phase space” ( N , d N d t ) Equation (4.1) is
described by Figure 4.2. (Remember that N < 0 representsin this model the population size and
therefore N < 0 is meaningless.)
Figure 4.2 stability of the steady states N = 1, N = 3. Arrows indicate the evolution of perturbations
from these states
We see from this figure that if N < 3 then d N d t < 0 and hence N is a decreasing function. On the
other hand if N > 3 then d N d t > 0 and N will increase (these facts are indicated by arrows in the
diagram). This is characteristic behavior of unstable steady state at N = 3. Similarly for the steady
state N = 1, d N d t > 0 if N < 1 (N is increasing with time) and d N d t < 0 if 1 < N < 3 (N is
decreasing with time). Hence N = 1 is stable. We point out that these results about the steady states of
this model were obtained without reference to the actual solutions of the model Equation (4.1).
Exercises
1. Solve the following equations in order to determine whether each of the steady states is
stable or unstable.
2. d N d t = a N - b N 2 , a, b > 0, N > 0
3. d N d t = - a N + b N 2 , a, b > 0, N > 0
4. d N d t = a N + b N 2 , a, b > 0 - ∞ < N < ∞
5. Use phase space techniques to deduce the stability of the steady states in Ex 1.
6. Use phase space techniques to discuss the stability of the steady states for
7. d N d t = α N + β N 2 + γ N 3 , β 2 - 4aγ > 0.
8. d N d t = ( N - α ) ( N - β ) ( N - γ ) . What happens when α = β = γ or α = β.
9. 4. A model of two societies M, N where M exploits N was suggested by May and Noy‐
Meir. According to this model if M, N are the population sizes of the two populations
respectively, then
dNft=aN(1-N/A)-BMN21+N2
where a, A, B M are constants. In this equation the first term represents the natural growth of society
N while the second represents the loss to this society due to the encounter between the two societies.
Discuss the steady states of this model and their stability when M is small, moderate, and large.
Hint: To find the steady states plot the two terms in the model equation separately. The steady states
are represented by the intersection of these two curves.
Figure 4.4 Regions in phase plane of (Equations 12.1) - (12.2) where d S d t is positive and negative
Figure 4.5 Stability of the steady state represented by the intersection of the lines d F d t = d S d t = 0
Figure 4.6 lntegral Curves of the Spring-mass system without friction
Definition 4.3.1
A system of ordinary differential equations
(4.15)
dxdt=F(x,t)
where x = ( x 1 ( t ) , … , x n ( t ) ) and
(4.16)
F(x,t)=f1(x,t)⋮ fn(x,t),x∈ Rn
is called autonomous if F(x, t) = F(x), i.e., the independent variable t does not appear explicitly in
F.
We observe that a non-autonomous system is equivalent to an autonomous system with an additional
equation. In fact if we define
(4.17)
z=xt,G(z)=F(x,t)1
then
(4.18)
dzdt=G(z)
is equivalent to (Equation 4.15)
In the following we consider only autonomous systems. Such systems are referred to as dynamical
systems.
Definition 4.3.2
The phase space of the system
(4.19)
dxdt=F(x)=f1(x)⋮ fn(x),x∈ Rn
is the space R 2n with coordinates ( x 1 , … , x n , x ˙ 1 , … x ˙ n ) ,
where x ˙ = d x i d t . Thus each point in phase space represents a unique state of the system. For a
system of particles satisfying Newton’s second law the phase space consists of all possible values of
the momentum and position variables.
Example 4.3.1
The equations of motion for a point particle of mass under the influence of an external force
F = F(x) is
(4.20)
md2xdt2=F(x),x∈ R3
which is equivalent to
(4.21)
dxdt=y,mdvdt=F(x).
Hence the phase space of Equation (4⋅ 20) is R 6 consisting of the points ( x , v ) .
Definition 4.3.3
A steady state(≡ critical point or equilibrium state) of the system (4⋅ 19) is a point x 0 such that
F(x 0) = 0.
We observe that a steady state can be isolated; i.e., there exists a neighborhood of x 0 which contains
no other steady state or it might be part of a continuous set of critical points.
Example 4.3.2
The system
(4.22)
dxdt=a1x+b1y,dydt=a2x+b2y,
has a continuous set of steady states when a 1 b 2 - b 1 a 2 = 0. In fact the steady state conditions d x
d t = d y d t = 0 lead to
(4.23)
a1x+b1y=0,a2x+b2y=0,
and these equations are linearly dependent (that is they represent the same equation) when a 1 b
2 - b 1 a 2 = 0. It follows then that any point on the line y = - a 1 x b 1 , b 1 ≠ 0 is a steady state of
the system. (When b 1 = 0, any point on the line x = 0 is a steady state.)
On the other hand this analysis shows that (0, 0) is an isolated steady state of the system represented
by Equation(4.22) when a 1 b 2 - b 1 a 2 ≠ 0, since then the coefficient matrix of (Equation 4.23) is
invertible.
In general it is rather difficult to analyze the stability of a system with a continuous set of steady
states. Therefore, in the following we restrict our discussion to systems with isolated critical points.
The following is a generalization of this example.
Theorem 4.3.1
Let F(x) be analytic and x 0 a steady state of the system represented by Equation (4 ⋅ 19). A
sufficient condition for x 0 to be isolated is that the Jacobian matrix of F(x) at x 0
(4.24)
J(x0)=[∂fi∂xj](x1)
is nonsingular.
Proof
Since F(x) is analytic, it has a Taylor series expansion that converges to F(x) . This Taylor expansion
can be used to approximate this function around the critical point x 0.
(4.25)
F(x)=F(x0)+J(x0)(x-x0)+O(|x-x02)=J(x0)(x-x0)+O(|x-x0|2).
Therefore, if another steady state exists in every neighborhood of x 0, then in the vicinity of x 0 all
higher order terms in Equation (4.25) become negligible in comparison to the linear term, and the
steady state must satisfy
(4.26)
J(x0)(x-x0)=0.
But J(x 0) is nonsingular which implies that the only solution of this system is x = x 0. Hence x 0 is
isolated. Note, however, that this theorem sets only a sufficient) but not necessary condition for x 0 to
be isolated.
Example 4.3.3
The only critical point of the system
(4.27)
dxdt=y2,dydt=x2
is (0, 0) although J(0, 0) is singular.
When a steady state is isolated, it is convenient, for conceptual and practical reasons, to translate the
steady state under consideration to the origin by the transformation w = x - x 0. The equations of the
system will become,
(4.28)
dwdt=G(w)
where G(w) = F(w + x 0)
Example 4.3.4
Consider the system
(4.29)
dxdt=4-xy,dydt=4x-y3
whose steady states are ( 2 , 2 ) and ( - 2 , - 2 ) . To translate ( 2 , 2 ) to the origin we perform the
transformation
(4.30)
u=x-2,v=y-2
and the resulting form of the system given by Equation (4 ⋅ 29), is
dudt=-2u-2v-uv
(4.31)
d v d t = 4 u - 12 v - 6 v 2 - v 3
Similarly if we wish to consider the steady state ( - 2 , - 2 ) we perform the translation
(4.32)
u=x+2,v=p+2
The resulting form of the system in Equation (4 ⋅ 29) is
dudt=2u+2v-uv
(4.33)
d v d t = 4 u - 12 v + 6 v 2 - v 3 .
Definition 4.3.4
A trajectory ( ≡ o r b i t , p a t h ) of the system given by Equation (4 ⋅ 19) is a solution of this
system x = x(t) subject to an initial condition x(0) = c.
Sometimes such a trajectory is written as x ( t , c ) .
Definition 4.3.5
Let x 0 be a steady state of the system, Equation (4 ⋅ 19) . We say that
1. x 0 is stable if for any given ɛ > 0 there exists a δ > 0 so that whenever |x(0) - x 0| < δ then
|x(t) - x 0| < ɛ for all t > 0.
2. x 0 is asymptotically stable if it satisfies the condition for being stable and in addition
|x(0) - x 0| < δ implies that
(4.34)
lim t → ∞ | x ( t ) - x 0 | = 0 .
Thus, x 0 is stable if whenever the initial state of the system is near the steady state then its trajectory
will remain in its vicinity for all times. On the other hand if these trajectories approach x 0 as t → ∞
then we say that x 0 is asymptotically stable.
Definition 4.3.6
A steady state that is neither stable nor asymptotically stable is called unstable.
Example 4.3.5
The differential equation governing the motion of the linear pendulum (without friction) which
was presented in Chapter 2, is
d2θdt2+ω2θ.
The general solution of this equation is
θ = A cos ( ω t + ϕ )
where A and φ are integration constants which are determined by the initial conditions. A small
perturbation from the steady state θ = 0, θ ˙ = 0 will take the system to a new (time dependent)
state but the “distance” between the equilibrium state and the new state will remain bounded and
does not go to 0 as time goes by. Hence the steady state is stable. On the other hand if we add
friction to this system the amplitude of the oscillations will decay to zero as t → ∞ and the
equilibrium state of the system becomes asymptotically stable.
Definition 4.3.7
The integral curves of the system represented by Equation (4 ⋅ 19) are the solutions of the system
(4.35)
dx1f1=…=dxnfn.
Thus the integral curves of Equation (4 ⋅ 19) are the trajectories of this system parametrized in
terms of the x i ′ s rather than in terms of the “extraneous” variable t.
Example 4.3.6
Find the trajectories and integral curves of
(4.36)
dxdt=αx,dydt=βy.
Solution 4.3.1 The trajectories of the system are
(4.37)
x=C1eαt,y=C2eβt
where C 1, C 2 are arbitrary constants. The corresponding integral curves are y α = Cx β where C is
a constant.
Example 4.3.7
Find the integral curves of the equations of the spring-mass system(with no friction).
Solution 4.3.2 The equation of motion for the spring mass system for small displacements from the
steady state is
(4.38)
d2xdt2+kx=0.
This is equivalent to the system
dxdt=v
(4.39)
dvdt=-kx.
The integral curves of this system satisfy
(4.40)
dxdv=v-kx
and hence
kx2+v2=C2
where C 2 is a constant. We conclude, therefore, that the integral curves of the spring‐mass system
are ellipses (see Figure 4.9). It follows from this result that the steady state (0, 0) of the system
given by Equation (4 ⋅ 39) is stable but not asymptotically stable.This result is obvious from a
physical point of view since the system contains no friction to damp the motion.
Exercises
1. Find the form of the following systems when each of their steady states is translated to the
origin.
dxdt=x( x- 2) ( y- 2) ,dydt=( x+2) ( y- 1) 2
d x d t = sin 2 π x cos 2 π y , d y d t = ( sin π y ) 2 - 1
Find the trajectories, integral curves and stability of the critical point at the origin for the
following systems
dxdt=- x,dydt=2x- y
dxdt=- 3x+y,dydt=x- 3y
d xd t= nz- ky, d yd t= kx- mz, d zd t= my- nx
Find the integral curves and stability of the critical points for the following systems.
ẍ=x3
m ẍ + b x ˙ + k x = 0 , m, b, k > 0. What happens if b < 0?
θ ¨ + ν 2 sin θ = 0
(nonlinear pendulum). What happens when |θ| is small and it is solution to approximate sin θ
by θ.
p( λ ) =a0λ n+…+a0
have non-positive real parts and all roots with zero real parts are simple.
Hint: Solve (Equation 4.54)
algorithm is given in Table 3.3.
For the following exercises find the critical points at which the system is almost linear and
discuss their stability.
2. d x d t = ( x - 1 ) ( x - 2 ) 2 ( y - 1 ) , d y d t = ( x - 1 ) 2 ( y - 2 )
3. d x d t = y ( 1 + cos x ) , d y d t = x ( 1 + sin y )
4. ẍ + β ( x 2 - 1 ) x ˙ + k x = 0 , b > 0. (Van der Pol eq.)
5. ẍ + β x ˙ - x + 2 x 3 = 0 (Duffin’s equation)
is positive definite on
Ω={ ( x,y) ;- b<x<b,- ∞<y<∞} .
where f(x) satisfies the assumptions of exercise 1, show that x = 0, x ˙ = 0 is a stable steady
state.
Hint: Let F ( x , x ˙ ) = 1 2 x ˙ 2 + α ∫ 0 x f ( t ) d t .
4. Apply the results of exercise 3 to the nonlinear pendulum
x ¨ + ω 2 sin x = 0 , ω 2 = g L .
5. Show that the following two systems are stable or asymptotically stable at 0
6. d u d t = - u - 2 u v 2 , d v d t = - v - v u 2
7. d u d t = - u 3 + 2 u v 2 , d v d t = - 2 u 2 v - 4 v 3 .
Hint: Consider a Liapounov function of the form
F( x,y) =au2+bv2,a,b>0.
8. For the system
dxdt=- ny- xf( x,y)
dydt=nx- yf( x,y) ,n=1,2,⋯
show that 0 is an asymptotically stable state when f ( x , y ) > 0 in some neighborhood of the
steady state and unstable if f ( x , y ) < 0 in some neighborhood of 0.
Hint: Note that this is a generalization of Example 4.6.5.
has limit cycles which correspond to the roots of the function h(r).
Hint: Use a polar representation of the system, (Equation 4.97).
3. Determine the periodic solutions and their stability for the system, (Equation 4.97) if
4. h(r) = r(r - 1)2(r - 3)(r - 4)
5. h(r) = r 2 - 4
6. h ( r ) = sin n r , n = 1, 2, . . .
7. Under what conditions on h(r), does the following system admit a limit cycle
(4.98)
dudt=2u+v- uh( r ) dvdt=- u+2v- vh( r )
is asymptotically stable if b > 0 and unstable if b < 0. Apply these results to the system,
(Equation 4.52), in Section 4.
Bibliography
[1] N.P. Bhatia, G.P. Szeg, Stability Theory of Dynamical Systems, Springer-Verlag, 2002
[2] R. Borrelli, C. Coleman, and W. Boyce, Differential Equations Laboratory Workbook: A Collection of Experiments, Explorations
and Modeling Projects for the Computer Wiley, NY, 1992.
[3] W.E. Boyce and R. Diprima, Elementary Differential Equations and Boundary Value Problems, 3rd edition, J. Wiley and Sons,
1977.
[4] Clive L. Dym - Stability Theory and Its Applications to Structural Mechanics, Dover, NY.
[5] M.W. Hirsch and S. Smale, Differential Equations, Dynamical Systems and Linear Algebra, Academic Press, 1974.
[6] R. K. Miller and A. N. Michel, Ordinary Differential Equations, Academic Press, 1982.
[7] J. Palis, Jr. and W. de Melo, Geometric Theory of Dynamical Systems, Springer-Verlag, 1982.
[8] A. Pillay - Geometric Stability Theory, Oxford University press 1996.
[9] N. Rouche, P. Habets and M. Laloy, Stability Theory by Liapounov's Direct Method, Springer-Verlag, 1977.
CHAPTER 5
CONTENTS
5.1 Introduction
5.2 Bifurcations of Co-Dimension One
5.2.1 Trans-critical Bifurcation
5.2.2 Saddle Point Bifurcation
5.2.3 Pitchfork Bifurcation
5.2.4 Subcritical Bifurcation (Hysteresis)
5.2.5 Hopf Bifurcation
5.3 Rossler Oscillator
5.4 Lorenz Equations
5.5 Nerve Models
5.6 Miscellaneous Topics
5.6.1 Dimension
5.6.2 Liapunov Exponents
5.7 Appendix A: Derivation Of Lorenz Equations
5.1 Introduction
In the previous chapter we discussed various methods to analyze the stability of the equilibrium states
of a dynamical system when the values of the system parameters are known and fixed. The objective
of bifurcation theory is to investigate what happens to the type, number, and stability of the steady
states as a result of a continuous change in some (or all) of the system parameters. If, as a result, the
system undergoes “sudden” change in its properties or behavior, we say that the system underwent a
bifurcation.
The motivation for such analysis stems from the fact that in many real life situations the values of
the model parameters are not known accurately or might actually be (very) slowly varying functions
of time which we approximate by constants to simplify the model Equations.
Here are some examples of “real life” bifurcations.
1. Phase transitions
If we cool down water, it will remain liquid until we reach 0 o C. If we continue the cooling
process below 0 o C, the water will undergo a “ phase transition” and turn to ice. Thus the “liquid
water system” undergoes a bifurcation at 0 o C. Similarly a bifurcation occurs when the water is
heated over 100C (at which point it turns to “gas”).
2. Earthquakes
The Earth’s continents “float” on the liquid core of the Earth. When these plates collide, stress
starts to build up, and one plate starts to glide over the other. As these stresses build up, a point
is reached where the plates break and we experience an earthquake. Thus an earthquake
represents a bifurcation in the plate’s structure.
3. Buckling of columns
If we increase the vertical load on a (vertical) column, a point will be reached where the
column will buckle and break (if the “load” is a building then the building will collapse.) Under
these circumstances the column undergoes a bifurcation.
4. Magnetic hysteresis
If we place a magnet in a weak magnetic field whose direction is opposite to its magnetization,
“nothing” will happen to the magnet. However, as we increase the field intensity, there will
come a point when all of a sudden the magnet will change its poles (the north magnet pole
becomes the southern one and vice versa). Thus a bifurcation occurs. If we reverse now the
direction of the field, another bifurcation will occur and the magnetization of the magnet will
return to its original state. These changes in the magnetization of the magnet are represented by
the “hysteresis loop” (see Fig. 5.5).
B.
(5.9)
x˙=λ+x2
The system has two steady states for x = ± - λ for λ < 0 and none for λ > 0 . In this case the steady
state x = - λ is unstable while x = - - λ is asymptotically stable (see Fig. 5.3).
and hence the steady state is asymptotically stable. Similarly one can show that x = - 1 is
asymptotically stable. While this analysis has been carried out only for λ = 0 , it is possible to show
that it is valid for other values of λ in the interval [ - 2 9 3 , 2 9 3 ] (as shown in Fig. 5.5). Thus on
this interval the system has one unstable state and two asymptotically stable states.
For values of λ which are outside the interval [ - 2 9 3 , 2 9 3 ] , it is straight- forward to show that
the corresponding steady state is asymptotically stable. For example, if λ = 6 , then the corresponding
steady state’s x = 2. Applying the transformation to local form Equation (5.2) to (Equation 5.14), viz.
μ=λ-6,y=x-2
we obtain
y = μ - 11 y - 6 y 2 - y 3
The linearized form of this Equation at μ = 0 is y ˙ = - 11 y which shows that the state y = 0 (viz.
x = 2) is asymptotically stable.
If initially λ < - 2 9 3 , then the system must be in the lower (stable) state. When λ becomes larger
than - 2 9 3 , the system acquires a second stable steady state but since the lower steady state is
stable, it will remain in its original state. However, when λ becomes larger than 2 9 3 , the lower
steady state ceases to exist and the system will “jump” to the upper stable steady state and will
remain there for all λ > 2 9 3 . A similar picture unfolds when λ is decreasing.
If λ > 2 9 3 , the system will be in the upper steady state and will remain in this state until λ = - 2 9
3 when this state ceases to exist. The actual state of the system is described therefore by the hysteresis
diagram (Fig. 5.5).
Figure 5.6 Hopf bifurcation: a steady state destabilizes as two conjugate eigenvalues move into the
the right hand side of the complex plane
Figure 5.11 Trajectory of the Rossler oscillator starting near x 1 with c = 8.2
Figure 5.12 Same trajectory in 3D
5.6.1 Dimension
We were all taught that a point has 0-dimension, a line is an entity of one dimension, a plane is two-
dimensional, and that “real space” has three dimensions.
We now try to formalize these intuitive concepts.
Consider the unit interval [0, 1 ] in R and consider measuring sticks of length ε = 1 2 n , n = 1,2,…
. The number M(ɛ) of such sticks that is needed to cover the interval [0, 1] is 2 n . Hence
(5.68)
D = lim ε → 0 ln M ( ε ) ln ( 1 ε ) = 1 .
(Here we dropped the requirement ε = 1 2 n ). Thus the expression, (Equation 5.68), yields a number
that corresponds to our intuitive notions about the dimension of a line.
Now consider the unit square [0, 1 ] × [0, 1 ] in the plane. Instead of a measuring stick we need
here a “ measuring square” with sides of 1 2 n . When n = 1
we need 4 such squares to cover [0, 1] × [0, 1]. For n = 2 we need 16 squares, and for ε = 1 2 n we
need 22n squares. Hence
(5.69)
D = lim ε → ∞ ln M ( ε ) ln ( 1 ε ) = 2 .
In three dimensions we consider the unit cube with “measuring cubes” whose sides are of length ε = 1
2 n . With n = 1 we need 8 such cubes to cover the unit cube. With n = 2 we need 64 cubes, etc.
Hence
(5.70)
D = lim ln M ( ε ) ln ( 1 ε ) = 3 .
Thus it seems appropriate to define the dimension of a metric space as
(5.71)
D = lim ε → 0 ln M ( ε ) ln ( 1 ε )
where M(ɛ) is the number of “cubes” with side ɛ that are needed to cover the unit cube in such a
space. The dimension D defined by (Equation 5.71) is called the “Hausdorff dimension” of the space.
We now apply this definition to an “exotic” set and compute its dimension.
Example 5.6.1 Cantor Set
To define the Cantor set we consider the interval I = [0, 1]. We now define
(5.72)
C1=I-(13,23)=[0,13]∪ [23,1],
i.e. C 1 is the subset that is obtained from the unit interval by removing the middle (open) 1 3 of I.
Now define C 2 as
(5.73)
C2=C1-(19,29)-79,89,
i.e. we remove again the middle 1 3 of each of the subintervals of C 1.
Similarly we define
(5.74)
C 3 = C 2 - ( 1 27 , 2 27 ) - ( 7 27 , 8 27 ) - ( 19 27 , 20 27 ) - ( 25 27 , 26 27 )
and so on. The Cantor set is now defined as the set remaining at the “end” of such interval divisions
and subtractions, i.e. (See Fig. 5.16)
(5.75)
C=∩ i=1∞Cn.
Figure 5.16 Cantor set
To show that C is not empty (in fact it is uncountable) we use the number representation with base
3. That is, a number 0 ≤ x ≤ 1 is represented by
(5.76)
x=.a1a2…,0≤ai≤2.
(That i s , t h e a i a r e integers which can take the values 0, 1, 2) where the a i s are determined by
the relation
(5.77)
x=∑n=1∞an3n.
Remark 5.6.1 The decimal and binary representation of a number 0 ≤ x ≤ 1 is a short hand notation
for
(5.78)
x = ∑ n = 1 ∞ α n 10 n , x = ∑ n = 1 ∞ β n 2 n
respectively where α i ε { 0 , 1 … 9 } , β i ɛ{0, 1 }.
It is now clear from its construction that the Cantor set consists of all numbers in base 3 whose
representation contains only the digits {0, 2 }; i.e. a i ≠ 1 for all i.
What is the Hausdorff dimension of the Cantor set?
To answer this question we consider a yard stick of length ε = 1 3 . It is then obvious that M ( 1 3 )
= 2 . Similarly for ε = 1 9 we have M ( 1 3 2 ) = 2 2 . In general we then have M ( 1 3 n ) = 2 n and
hence
(5.79)
D = lim ln M ( ε ) ln ( 1 ε ) = ln 2 ln 3 .
Thus the Cantor set is a set with non-integer Hausdorff dimension.
Example 5.6.2 Sierpinski Triangle
To construct this set we start with an equilateral triangle S 0 with sides of length one. In the first
iteration we split it into four equal-sided triangles (as shown in Fig. 5.17) and remove the one in
the center. This yields S 1. By iterating this procedure on each of the three remaining triangles we
obtain S 2 and so on. The Sierpinski triangle is the limit set that remains after all these divisions
and removals.
Figure 5.17 First two iterations toward the creation of the Sierpinski Triangle
S = lim n → ∞ S n .
To find the Hausdorff dimension of S we note that for S 0 we have one triangle and need a
measuring stick with length ɛ = 1. For S 1 we have three triangles and ε = 1 2 . For S 2 we have nine
triangle and ε = 1 4 . Hence
D = lim ε → 0 ln M ( ε ) ln ( 1 ε ) ≡ lim n → ∞ ln 3 n ln 2 n = ln 3 ln 2 .
Thus the dimension of this “triangle” is greater than one but less than two. To compute the area of
S we observe that S 0 has area 3 4 . S 1 has an area of 3 4 . 3 4 (since 1 4 th of the area has been
removed). S 2 has an area of ( 3 4 3 4 ) · 3 4 (since we removed once again 1 4 of the area of S 1). In
general we then have
areaofSn=34(34)n
and hence
area of S = lim n → ∞ (area of S n ) = 0.
Perturbations
CONTENTS
6.1 Introduction
6.3 Regular Perturbations
6.4 Singular Perturbations
6.5 Boundary Layers
6.1 INTRODUCTION
When we model a complex system, we create at first a “prototype model.” This model takes into
account the “major features” of the system at hand and ignores other aspects which might have “small
impact” on the system evolution. In other words the prototype model reduces the system to its “bare
bones.” In many cases these prototype models lead to equations with closed form solutions, viz.
solutions in terms of analytical formulas. However, when refinements of the prototype model are
needed, new (nonlinear) terms have to be added, and in many cases the resulting model equations
have no closed form solutions. At this stage one realizes that the additional terms in the refined model
might be “small” when compared to original terms in the prototype model and attempts to take
advantage of this fact. The essence of perturbation theory and techniques is to develop methods that
yield at least approximate solutions for these refined models. As a first step in this approach one has
to rewrite the model equations in non-dimensional form. That is, the equations have to be rewritten in
a form that is independent of the physical units that are being used. In this form one can see
immediately which terms are small when compared to other terms in the equations and thereby treat
them as a “perturbation.”
It might be argued that with wide availability of fast computers and appropriate software one can
solve these complicated (nonlinear) equations (of the refined model) numerically. However, in
general, it is not straightforward to gain insights about the evolution of the system from the numerical
solution. Furthermore, the numerical solution of a complex system of nonlinear differential equations
is always a challenging problem. Algorithms to solve such a system might not converge, and even if
they converge, they might yield incorrect solutions.
In the following we present some of the basic techniques of perturbations theory, but for a more
comprehensive treatment we refer the reader to books on this topic.
When we model a physical system, the variables and the value of the coefficients that appear in the
model equations depend on the physical units being used. For example the length unit might be
centimeter, meter, kilometer, light year (i.e. the distance that light travels in one Earth year) and so on.
It is obvious then that the use of different units will impact the value of the parameters and
coefficients that appear in the model equation. In order to overcome this issue one attempts to find
“characteristic values” for the variables that appear in the equations and combinations of these values
so that the equations are expressed in terms of “pure numbers” (independent of the physical units). We
illustrate this process by a few examples.
Example 6.2.1
Consider Newton’s second law
(6.1)
md2xdt2=F.
To rewrite this law in non‐dimensional form let us assume that the characteristic values of mass \,
length, and time in this equation are M, L, and T respectively. The force term F in this equation
has a dimension of m a s s · l e n g t h t i m e 2 . Therefore, we define the following dimensionless
quantities,
(6.2)
m¯=mM,x¯=xL,t¯=tT,F¯=FT2ML.
Then the left hand side of Equation (12.1) becomes
md2xdt2=Mm-d2(Lx¯)d(T2t¯)=MLT2m-d2X¯dt¯.
At the same time the right hand side of Equation (12.1) yields
F=MLT2F¯.
Therefore, Equation (12.1) takes the following form
(6.3)
m-d2x¯dt¯=F¯.
Although Equations (12.1), (12.3) look “similar, “ the quantities in (12.3) are dimensionless (i.e.
independent of the physical units).
To see why this non‐dimensionalization process is important we the discuss the following
example.
Example 6.2.2
Suppose we consider the following modification of Equation (12.1)
(6.4)
md2xdt2=F+α(dxdt)2.
The question naturally arises as to the impact of the additional term on the solution. Is it “small”
or “large” compared to the other terms of the equatio n ? To answer this question we introduce the
same non‐dimensional quantities as in (12.2), and the additional term in Equation (12.4) becomes
(dxdt)2=(d(Lx¯)d(Tt¯))2=L2T2(dx¯dt¯)2
Therefore, Equation (12.4) leads to
(6.5)
MLT2m-d2x¯dt¯=MLT2F¯+αL2T2(dx¯dt¯)2
Hence
(6.6)
m-d2X¯dt¯=F¯+αLM(dx¯dt¯)2.
It follows then that the size of the additional term in Equation (12.4) (as compared to other terms
in this equation) is not determined by α alone but by the ratio α ¯ = α L M . If α ¯ ≪ 1 , we can
treat the additional term in Equation (12.4) as a small perturbation. On the other hand if α ¯ ≈ 1 ,
such a treatment will be inappropriate.
Remark
Another way to look on this issue is to realize that α is not a dimensionless number since each of
the terms in Equation (12.4) must have the same dimension. TherefO re α must have the dimension
of m a s s l e n g t h . Hence the dimensionless form of this constant is α L M .
Additional examples of the non‐dimensionalization process are presented in Chapters 9 and .
Example 6.3.1
Find an approximate solution to the following equation,
(6.7)
dydt+ky+εy3=0,y(0)=2.
Here we assume that the equation is written already in non‐dimensional form and ɛ ≪ 1.
Solution 6.3.1
If ɛ = 0 the solution of Equation (12.23) is
(6.8)
y(t)=2e-kt.
To find approximate solution of (12.23) when ɛ ≠ 0 we write the desired solution y ( t , ε ) in the
form of a power series around ɛ = 0viz.
(6.9)
y(t,ε)=y0(t)+εy1(t)+ε2y2(t)+…
Substituting this expansion in Equation (12.23) and collecting terms with the same power of ɛ
yields,
(6.10)
[dy0dt+ky0]+ε[dy1dt+ky1+y03]+ε2[dy2dt+ky2+3y02y1]+…=0.
(6.11)
y0(0)+εy1(0)+ε2y2(0)+…=2.
Since ɛ is a parameter, terms with diffe rent powers of ɛ in Equation (12.26) must each equal to
zero. Using Equation (12.27) we obtain
y0(0)=2,yi(0)=0,i≠0.
Thus we obtain the following system of diffe rential equations
(6.12)
dy0dt+ky0=0,y0(0)=2,
(6.13)
dy1dt+ky1+y03=0,y1(0)=0,
(6.14)
dy2dt+ky2+3y02y1,y2(0)=0,
and so on. Observe that although the original Equation (12.23) is nonlinear the new system of
equations is actually linear and can be solved recursively. In fact the solution for y 0 is given by
Equation (12.24). Substituting this solution in Equation (9.21) we obtain
y1(t)=4k(e-2kt-1)e-kt
Similarly we can substitute the solutions for y 0 and y 1 in Equation (9.22) and solve for y 2 etc.
Hence to order ɛ the (approximate) solution of Equation(12.23) is
(6.15)
ya(t)=2e-kt+4εk(e-2kt-1)e-kt.
It is interesting to note that although, Equation (12.23) is nonlinear, it has an exact closed form
solution
(6.16)
ye(t)=2k-4ε+e2kt(4ε+k).
In Fig 6.1 we plotted the diffe rence y a - y e on the time interval [0, 5 ] for k = 1 and ɛ = 0.1.
Exercises
1. Use regular perturbations to find an approximate solution to order ɛ of the following equation
dydx+εy+3y2=0,y(0)=1.
2. The equation for the orbit of the planet Mercury around the Sun (in general relativity) can be
reduced to the following equation
d2udθ2+u=a(1+εu2)
where ɛ ≪ 1. Use first order perturbations to examine the impact of the term aɛu 2 on the period of the
orbit.
Hint: First solve this equation with ɛ = 0.
3. Use regular perturbations to find an approximate solution to order ɛ of the following initial value
problem
d2ydx2+4y-εy2=0,
To see the challenge that these types of problems present (from perturbation theory perspective) we
attempt, at first, to obtain an approximate solution via regular perturbations. Thus, if we write the
perturbation expansion
(6.36)
y(x)=y0(x)+εy1(x)+ε2y2(x)+…
and substitute in (Equation 6.35) we obtain the following system of equations:
(6.37)
dy0dx+ky0=0,y0(0)=0,y0(1)=1
(6.38)
dy1dx+ky1=-d2y0dx2,y1(0)=y1(1)=0
(6.39)
dy2dx+ky2=-d2y1dx2,y2(0)=y2(1)=0
etc. Thus although (Equation 6.35) is a second order equation with two boundary conditions, the
perturbation expansion yields a system of first order equations with two boundary conditions for
which no solution exists in general. In fact the general solution of (Equation 6.37) is
y0=Ce-kx,
but since y 0(0) = 0, it follows that C = 0 and, therefore, y 0(1) = 1 cannot be satisfied.
To see the origin of this peculiar behavior and how it can be ameliorated we note that the second
order derivative of the exact solution of (Equation 6.35) has a factor of
exp(-x2ε)ε
This observation shows that the term ε d 2 y d x 2 is small compared to d y d x and y for large x (e.g.,
x > 2 ε b . However, in a small interval around 0 it is of the same order as these two terms.
In view of this observation it is natural to divide the solution of our problem into two.
1. A solution in an inner region 0 ≥ x ≤ x 0 where x 0 > 0 is small.
2. A solution in an exterior region x 0 < x.
Observe that, at the present, the exact value of x 0 remains unspecified.
Denoting the exterior solution by v we note that in this region v, and its first and second order
derivatives have the same order of magnitude, and hence ε d 2 v d x 2 is negligible as compared to
the other terms in (Equation 6.35). Therefore, in the exterior region (Equation 6.35) can be
approximated correctly by a regular perturbation expansion,
(6.40)
dv0dx+kv0=0,v0(1)=1,
(6.41)
dv1dx+kv1=-d2v0dx2,v1(1)=0,
etc.
Observe that the point x = 0 is outside the “exterior region.” Therefore, we do not have to take into
account the boundary condition on the solution at this point.
From (Equations 6.40) and (6.41) we obtain
(6.42)
v 0 = exp [ - k ( x - 1 ) ]
(6.43)
v 1 = - k 2 ( x - 1 ) exp [ - k ( x - 1 ) ] ,
i.e., to first order in ɛ the outer solution is given by
(6.44)
v = [ 1 - ε k 2 ( x - 1 ) ] exp [ - k ( x - 1 ) ] .
To obtain a similar perturbation expansion in the inner region we shall perform a stretching
transformation on this region so that the derivatives of the solution in this region have the “same
magnitude.” This can be accomplished by a transformation of the form
(6.45)
z=εaX
where a is a parameter to be determined. Substituting this transformation in (Equation 6.35) and
denoting the inner solution by u it follows that
(6.46)
ε1+2ad2udz2+εadudz+ku=0,u(O)=0
From this equation it follows that the derivatives of u will have the same power of ɛ if 1 + 2a = a i.e.,
a = - 1. (Observe that due to this stretching transformation x = ɛ will correspond to z = 1.) Hence in
the inner region (Equation 6.35) takes the form
(6.47)
d2udz2+dudz+εku=0,u(O)=0.
Applying regular perturbation expansion to (Equation 6.47) yields
(6.48)
d2u0dz2+du0dz=0,u0(0)=0
(6.49)
d2u1dz2+du1dz=-ku0,u1(0)=0
and so on.
The solution of these equations to first order in ɛ is given by
(6.50)
u0=c1(1-e-z).
(6.51)
u1=k[c2-c1z-(c2+c1z)e-z].
Now that we have perturbation solutions for the inner and outer regions a “matching principal” must
be formulated so that the two solutions blend smoothly at the “edges” of the two regions. This will
help to determine the constants c 1, c 2 which appear in (Equations 6.50) and (6.51). To accomplish
this task several “principles” were formulated in the past. The first due to Prandtl stipulates that
(6.52)
lim x → 0 v ( x , ε ) = lim z → ∞ u ( z , ε ) .
Applying this principle to the zeroth order perturbation solutions (6.42) and (6.51) of our problem we
find that we must satisfy
(6.53)
lim x → 0 A e k ( 1 - x ) = lim z → ∞ c 1 ( 1 - e - z ) .
Hence c 1 = Ae k . However, it is easy to verify that this principle fails to determine the constant c 2 in
the first order perturbation expansion.
To overcome this problem M. Van Dyke formulated the following generalized matching principle
for higher order perturbation expansions. To apply this matching principle to the perturbation
expansions in the inner and outer regions up to order ɛ m ,
v(x)=v0(x)+εv1(x)+…+εmvm(x)
uz=u0z+εu1z+…+εmumz,
one must express v(x) in terms of z and u(z) in terms of x, then expand the resulting expressions in
powers of ɛ up to ɛ m and finally, match the coefficients of ɛ k , k = 0, . . ., m in the two expressions to
determine the redundant constants.
As an example we implement this principle to determine the constants c 1, c 2 in (Equations 6.50),
(6.51) by matching the inner and outer solutions to first order in ɛ.
Rewriting the outer solution (Equation 6.44) in terms of z and expanding in ɛ we obtain
(6.54)
v(z)=Aek[e-εkz+εk2(1-εz)e-εkz]
=Aek[1-εkz+εk2+O(ε2)].
Similarly for the inner solution we have
(6.55)
u(x)=c1(1-e-x/ε)+εk[c2-c1x/ε-
(c2+c1x/ε)e-x/ε]=c1(1-kx)+εkc2+O(ε2).
(Note that the expression e -x/ɛ for any fixed x > 0 converge to 0 faster than any powers of ɛ and,
therefore, can be neglected). It follows then,
(6.56)
c1(1-kx)+εkc2=Aek[(1-kx)+εk2].
This yields
(6.57)
c1=Aek,c2=Akek.
Exercises
For the following two differential equations use first order perturbations to find inner, outer, and
matched solutions. Solve these equations numerically.
Compare the numerical and analytic solutions on the same plot.
εd2ydx2+(1+ax)dydx+k2y=0,y(O)=0y(1)=1.
εd2ydx2+(1+ε)dydx+y2=0,y(O)=0y(1)=1.
CHAPTER 7
CONTENTS
7.1 The Heat (or Diffusion) Equation
7.1.1 Burger’s Equation
7.1.2 Similarity Solutions
7.1.3 Stephan Problem(s)
7.2 Modeling Wave Phenomena
7.2.1 Nonlinear Wave Equations
7.2.2 Riemann Invariants
7.3 Shallow Water Waves
7.3.1 Tsunamis
7.4 Uniform Transmission Line
7.5 The Potential (Or Laplace) Equation
7.5.1 Kirchoff Transformation
7.6 The Continuity Equation
7.7 Electromagnetism
7.7.1 Maxwell Equations
7.7.2 Electrostatic Fields
7.7.3 Multipole Expansion
7.7.4 Magnetostatic
7.7.5 Electromagnetic Waves
7.7.6 Electromagnetic Energy and Momentum
7.7.7 Electromagnetic Potential
In this chapter, we illustrate this modeling process by considering various systems that are modeled in
terms of partial differential equations. In particular, we concentrate on the heat, wave, and potential
equations that are important in many scientific and engineering applications.
Background
To build an acceptable model for this problem, an understanding of the concept of the flux and
basic laws of thermodynamics (and physics) is necessary. We present here a short review of these
pertinent ideas in one spatial dimension.
The Flux. Consider a flow of a certain physical quantity (such as mass, energy, heat, etc.). The flux
q(x,t) of this flow is defined as a vector in the direction of the flow [at (x, t)] whose length is given by
the amount of the quantity crossing a unit area (at x) normal to the flow in unit time; that is,
(7.1)
| q ( x , t ) | = △ S → 0 lim △ t → 0 Q u a n t i t y p a s s i n g t h r o u g h △ S i n t i m e [ t , t +
△ t]△ S△ t
where AS is a (small) surface area at x that is normal to the flow.
Thus, the approximate amount of the physical quantity passing through a
surface AS in time At is given by
(7.2)
Q(x,t,△ S,△ t)≅ |q(x,t)|△ S△ t
If AS is not normal to the flow, then it must be replaced by its projection in the direction normal to the
flow.
Figure 7.2 Flux-All the fluid in the tube will pass through S in time Δt
Example 7.1.1
Consider water flowing in a river with velocity v(x,t). To evaluate the flux of this flow at (x, t ) , we
consider a small surface element A S normal to v ( x,t ) . The amount of water flowing through A S
in time [ t,t + At] is given by the quantity present at t in a tube of base A S and height |v|At; that is
(7.3)
Q(x,t,△ S,△ t)≅ ρ|v|△ S△ t
where p is the mass density of the water. Hence,
(7.4)
q(x,t)=ρv
since the direction of the flow is given by v (see Fig. 7.2 ).
Basic Laws of Thermodynamics. A change AQ in the amount of heat in a body of mass m is
accompanied by a change A u in its (equilibrium) temperature. The relationship between these
changes is given by
(7.5)
△ Q=cm△ u
where c is the specific heat of the material of which the body is made, that is, the amount of heat
required to raise the temperature of a body of unit mass (made of the same material) by 1 degree.
In the following discussion, we assume that Q and u are normalized so that Q = cmu .
Remark About Units. In (Equation 7.3) (as in any equation that relates physical quantities), a
consistent set of units must be used. Thus, if the MKS system of units is used, then Q (energy) is
expressed in joules, mass in kilograms, u in degrees Kelvin (or Celsius), and c in joules/(kg.degK). In
this book we consistently use the MKS units unless otherwise noted.
Fourier Law of Heat Conduction . Heat is transported by diffusion in the direction opposite to the
temperature gradient and at a rate proportional to it. Thus, the heat flux q(x, t) is related to the
temperature gradient by
(7.6)
q(x,t)=-κgradu(x,t)=-κ(∂u(x,t)∂x,∂u(x,t)∂y,∂u(x.t)∂z)
where κ is the thermal conductivity of the material. [From (Equation 7.6) we infer that its units are
joules/(m.sec. degK).]
Remember that the gradient of a function gives the direction in which the function increases most
rapidly while in the direction opposite to it the function decreases most rapidly. Thus, a restatement
of the Fourier law is that heat flows in the direction in which the temperature decreases most rapidly
(and this is the reason for the minus sign in (Equation 7.6). Note that by convention any constant that
appears in a physical law is assumed to be positive).
Principle of Energy Conservation. Because heat is a form of energy, it must be conserved. Hence,
the rate of change in the amount of heat in a body must equal the rate at which heat is flowing in less
the rate at which it is flowing out (we assume that no heat is generated by the body).
Approximations and Idealizations
1. Since we assumed that the material of the rod that we are considering is homogeneous, it
follows that c, k, and ρ (the material density kg/m 3 ) are independent of the position x.
However, for the purpose of constructing a prototype model we further assume that they are
also independent of the temperature u .
2. We assume that the length of the rod remains constant in spite of the changes in its
temperature.
3. We assume that the rod is perfectly insulated along its lateral surface (idealization). Hence,
heat can flow only in the horizontal direction, since a vertical flow will lead to heat
accumulation along the edges, which is forbidden by the Fourier law of conduction.
Therefore, we infer that the temperature at every point on a vertical cross section of the rod
must be the same. Thus, the temperature u depends only on x and t; that is, u = u(x, t).
4. For definiteness, we assume that heat flows in the rod from left to right, which requires the
left side to be warmer than the right.
Modeling
A possible approach to model the system under consideration is to use the atomic and crystal
structure of the material of the rod and build a model for the heat conduction using these microscopic
variables. However, because this approach will lead to a very complex set of equations, it is not
useful in our context.
Thus, we present in the following two methods to derive the macroscopic heat equation. We use the
term macroscopic since we are using macroscopic variables such as u, c, k, and so on to model the
system.
Infinitesimal Approach
In this method, we consider an infinitesimal element of the rod between x and x + Δχ and write the
equation for the energy conservation in it.
Thus, since the volume of the element is ΑΔχ, its mass Δη is given by ρΑ Δ χ (see Fig. 7.3). The
amount of heat in this element at time t is or
(7.7)
Q(x,t,△ x)=c△ mu(x,t)
Figure 7.3 Heat flux through infinitesimal part of the rod.
or
(7.8)
Q(x,t,△ x)=cρAu(x,t)△ x.
The rate of change in Q is therefore given by
(7.9)
dQdt=cρA∂u∂t△ x.
By the principle of energy conservation, this rate of change must equal the rate at which heat is
flowing in less the rate at which it is flowing out. Hence,
(7.10)
dQdt=q(x,t)·A-q(x+△ x,t)·A.
Replacing d Q d t by
(7.11)
cρA∂u∂t△ x
we have
(7.12)
cρA∂u∂t=-Aq(x+△ x,t)-q(x,t)△ x.
Letting Δχ 0, we obtain
(7.13)
cρ∂u∂t=-∂q∂x.
From (7.6), the Fourier law of heat conduction in one dimension yields q =- κ( ∂ u/ ∂ x) (since u is a
function of x and t only!) and, therefore,
(7.14)
cρ∂u∂t=κ∂2u∂x2
or
(7.15)
1k∂u∂t=∂2u∂x2
where k−1 = cp/κ is called the thermal diffusivity. (Equation 7.15) is called the heat (or diffusion)
equation in one (space) dimension.
Integral Approach
In this method, we consider a finite section of the rod between a and b and use the principle of
energy conservation to write an equation for the heat balance in this segment.
Since the amount of heat in an infinitesimal section of the rod between x and x + Δχ is given by
(Equation 7.8), the total amount of heat in the section [a, b] is given by the integral of the expression
(7.16)
Q(t,a,b)=∫abcρAu(x,t)dx.
The rate of change in this quantity is therefore given by
(7.17)
dQdt=∫abcρA∂u∂tdx.
By the principle of energy conservation, dQ/dt must equal the rate at which heat enters the section
less the rate at which it leaves it. Thus,
(7.18)
dQdt=Aq(a,t)-Aq(b,t).
By the fundamental theorem of calculus, this equation can be rewritten as
(7.19)
dQdt=-∫abA∂q∂xdx
from which it follows that
(7.20)
∫abcρA∂u∂tdx=-∫abA∂q∂xdx.
By the Fourier law of heat conduction,
∫ab(cρA∂u∂t-κA∂2u∂x2)dx=0
(7.21)
∫ab(cρA∂u∂t-κA∂2u∂x2)dx=0.
But since a and b are arbitrary, Equation (17.21) implies that the integrand in this equation must also
be zero. Hence,
(7.22)
cρ∂u∂t-κ∂2u∂x2=0
which is the same equation we derived using the infinitesimal approach.
Remark 7.1.1
Although the infinitesimal and integral approaches must always (if applied correctly) yield the
same result, from a conceptual modeling point of view one might be superior to the other in a
given context.
To illustrate the process of model compounding, we now present the derivation of the heat equation
in two dimensions.
Example 7.1.2
Derive the heat equation for a thin homogeneous plate with constant cross section (height) h.
Assume that the plate is perfectly insulated on the top and bottom.
Solution 7.1.1 Since the plate is thin and perfectly insulated,
(7.23)
u=u(x,y,t).
To derive a model for u, we use the infinitesimal approach and consider a small rectangular element
that is located at a point (x, y) in the plate (see Fig. 7.4).
Background
In general, an in-depth treatment of wave phenomena requires considering the elastic properties of
matter and leads to a complicated set of equations. To overcome this difficulty, we make the
following simplifying approximations and idealizations so that a prototype model can be constructed
by applying only Newton’s second law F = ma (i.e., the force equals the mass multiplied by the
acceleration) to the system under study.
Approximations and Idealizations
1. The string is rigidly attached at its end points.
2. The string vibrates in one plane.
3. No external forces act on the string (prototype model).
4. The string does not suffer from damping forces (prototype model).
5. The string is homogeneous. In particular, this implies that the linear density ρ (i.e., the mass
per unit length) of the string is constant.
6. The deflection u of the string from equilibrium and its slope are always small.
Consequently, we are able to assume a point on the string moves only in the vertical
direction.
7. The tension force in the string is always tangential to it. This is usually expressed by saying
that the string is assumed to be perfectly flexible.
Modeling
Consider a small segment of the string between x and x + Δχ as shown in Fig. 7.6. Before we can
apply Newton’s Second Law to the motion of this segment, we must make the following observations:
1. By approximation 6, the segment is not moving in the horizontal direction. Therefore if we
denote by T(x,t) the tension in the string, the horizontal components of T(x,t) at x and x + Δχ
must be equal (see Fig 7.6).
Figure 7.6 Balance of forces on a small section of the string
(7.40)
T ( x ) cos α = T ( x + △ x ) cos β = R
1. The mass of the segment [ x , x + △ x ] is given by ρ ▵ s. Where ▵ s is the length of this
segment in the deformed state of the string. However, since we are considering only small
deflections, |u| 1, it follows that
△ s≈△ x.
1. The acceleration of the segment in the vertical direction is given by
(∂2u/∂t2).
1. The sum of the vertical forces acting on the segment (see Fig. 7.6)is
T ( x + △ x ) sin β - T ( x ) sin α .
Combining all these observations and approximations, we infer from Newton’s second law that
(7.41)
ρ △ x ∂ 2 u ∂ t 2 = T ( x + △ x ) sin β - T ( x ) sin α .
Using (7.40) to elliminate T form this equation yields
(7.42)
ρ△ x∂2u∂t2=R(tanβ-tanα)
However since T is always tangential to the string
(7.43)
tan α = ∂ u ∂ x ( x , t )
(7.44)
tan β = ∂ u ∂ x ( x + △ x , t ) .
Substituting (Equations 7.43) and (7.44) in (Equation 7.42) and dividing by ▵ x, we obtain
(7.45)
ρ∂2u∂t2=R∂u∂x(x+△ x,t)-∂u∂x(x,t)△ x.
Letting ▵ x → 0, we finally obtain
(7.46)
1c2∂2u∂t2=∂2u∂x2
where c 2 = R/ρ.
(Equation 7.46) is called the wave equation in one dimension.
Remark 7.2.1 From Fig. 7.6, we can infer that the sum of the horizon‐ tal forces acting on the
string segment is T ( cos β - cos α ) ≠ 0 . Hence, the segment must have an acceleration in the
horizontal direction, which contradicts approximation 6b. However, since α and β are small, we
can argue that T ( cos β - cos α ) is negligible.
Compounding
Example 7.2.1
Derive a model equation for very small vibrations of a vertically suspended chain whose length is
L and whose mass density per unit length p is constant.
Approximations
1. Since the amplitude u of the vibrations is small, we assume that a point on the chain does
not change its x-coordinate (see Fig 7.7).
2. The tension T(x,t) in the chain cannot be assumed to be constant in the context of this
problem. In fact, in the equilibrium (vertical) position of the chain,
(7.47)
T(x)=ρg(L-x)
In the following model, we assume that (Equation 7.47) gives an accept‐ able approximation for the
tension in the vibrating chain when |u| ≪ 1 and | ∂ u ∂ t | ≪ 1 .
1. Other approximations and idealizations of the prototype model remain intact.
Modeling
To construct a mathematical model, we once again consider a small section of chain between [ x , x
+ △ x ] . Applying Newton’s second law in the horizontal direction to such a section (see Fig. 7.8),
we obtain
(7.48)
ρ △ x ∂ 2 u ∂ t 2 = T ( x + △ x ) sin β - T ( x ) sin α .
But since α and β are small, we can once again use the approximations given by (Equations 7.43)
and (7.44). Hence,
(7.49)
ρ∂2u∂t2=1△ x[T(x+△ x)∂u∂x(x+△ x,t)-T(x)∂u∂x(x,t)].
Letting ▵ x → 0, it follows then that
(7.50)
ρ∂2u∂t2=∂∂x[T(x)∂u∂x(x,t)].
Substituting (Equation 7.47) for T(x), we finally obtain
(7.51)
∂2u∂t2=g∂∂x[(L-x)∂u∂x].
Other examples of physical systems describing wave phenomena will be discussed in Sections 4 and
5.
Hint. Let p be the mass density of the bar and A its cross section. The mass of the small section is
pAAx, and its acceleration is d 2 u/dt 2 . The stress difference across the section is EA[u x (x + Ax,t)
— ux(x,t)]. Note that c = y/E/p is called the speed of sound in the medium.
1. Derive a model equation for the vibrations of a membrane whose edges are fixed. Make
similar assumptions as for the vibrating string.
Hint. Consider a small rectangular area and apply the same analysis as for the string in the x and y
directions.
1. The pressure p and the mass flow rate u of a fluid flowing in a long pipe are related
approximately by the equations
∂p∂t=c∂u∂x
∂u∂p
∂t¯=∂x¯
where c is the compressibility of the fluid. Show that both p nd u satisfy the wave equation in one
dimension.
1. Show that
u(x,t)=f(x-ct)+g(x+ct)
is a solution of the wave equation in one dimension when f and g are any “smooth functions.”
1. Derive a model equation for the small vibrations of a string that is rotating around the x-axis
at a constant angular velocity ω. Assume that at each moment all the points of the string are
one plane.
2. A mass m is attached to the end of a suspended chain of length L and linear density p .
Derive a model equation for the small vibrations of this system.
3. The results of Exercise 4 can be applied to the air vibrations in an organ pipe. However,
since it is not “easy” to follow the position of air molecules, it is more natural to derive an
equation for the pressure in the pipe. If it is known that the pressure is proportional to du/dx
, show that it must satisfy the wave equation.
7.3.1 Tsunamis
Waves on the oceans are generated by various causes. First, there are waves generated by winds
blowing on the ocean surface. Then, there are waves due to tidal forces exerted by the gravitational
interaction with the Sun and Moon systems. Finally, there are those that are generated as a result of
earthquakes or other natural catastrophes. This last category of waves is referred to as tsunamis.
Tsunamis are characterized by long wavelength λ and period ω. Typically, the wavelength can
range from few to hundreds of kilometers. Since the ocean depth D on Earth is less than llkm, it
follows that for these waves we always have λ / D ≫ 1 . Accordingly, tsunamis can be treated as
shallow water waves. As the wave approaches, land D decreases and λ / D becomes very large.
Thus, decreasing depth leads to a higher concentration of the energy.
In the deep ocean, we can approximate the horizontal and vertical dis‐ placements of a particle at (
x , y ) due to the wave by
X = A sin ( k x + ω ( k ) t ) , h = - B cos ( k x + ω ( k ) t )
where A is a constant and B = Aky. The factor ky has been added in this model of the tsunami wave
for several reasons. To begin with, k = 2 π / λ , and it follows that at the ocean surface the vertical
amplitude of this wave is 2 π A D / λ . However, D / λ ≪ 1 , which implies that (as expected) the
vertical amplitude of the wave is much smaller than the horizontal amplitude. In addition, the vertical
displacement is 0 at the ocean bottom. Furthermore, this form of the wave satisfies the continuity
equation div u = 0, u = ( u , v ) . In fact, the velocity components of the wave are
u = A ω cos ( k x + ω ( k ) t ) , v = A k y ω sin ( k x + ω ( k ) t )
and therefore
ux+vy=0.
Energy conservation implies then that ω 2 = (gD)k 2. (Such a relation that expresses the frequency as a
function of the wave number is called ((a dispersion relation.”) The group velocity is therefore
vg=|dωdk|=gD.
Observe that for this dispersion relation, the phase velocity v p = | ω k | is equal also to g D . From
this expression, we infer that in the deep ocean with D ≈ 4km, we have v g ≈ 200m/sec. Thus, a
tsunami can propagate very fast in the open ocean. However, as the water depth decreases, the
tsunami slows down. At the same time the tsunami’s energy flux, which is the sum of its kinetic and
potential energies, remains almost constant. Consequently, as the tsunami’s speed diminishes as it
travels into shallower water, its height grows. Because of this effect, a tsunami, imperceptible at sea,
may grow to be several meters or more in height near the coast.
Thus, assuming no dissipation due to friction or turbulence, the tsunami energy flux is proportional
to B 2 v g or B 2 D . It follows then that constant energy flow requires that the tsunami amplitude B is
proportional to D -1/4. Suppose now that at a depth of 4km the tsunami’s vertical amplitude is lm (a
“normal wave”), then at depth of 2m (near the shore) the wave height will be approximately 6.7m
Exercises
1. If D is a constant and u, h, and their derivatives are small (so that nonlinear terms in these
quantities can be neglected), show that h ( x , t ) must satisfy the wave equation
htt=(gD)hxx.
1. Under the same conditions as in the previous exercise, what is the equation that is satisfied
by u(x,t)?
2. Repeat Exercises 1 and 2 when D is a linear function of x; that is,
D=ax+ba>0,b<0.
1. Modify (Equation 7.71) to include the action of a damping force (in the x direction) that is
proportional to u.
Figure 7.12 Equivalent circuit for a small section of the transmission line
Applying Kirchoff’s second law between A and D, (see Fig. 7.12) we obtain
(7.76)
e(x,t)-e(x+△ x,t)=R△ xi(x,t)+L△ x∂i(x,t)∂t.
Similarly, applying Kirchoff’s first law at the node B, we have
(7.77)
i(x,t)-i(x+△ x,t)=C△ x∂e(x+△ x,t)∂t+G△ xe(x+△ x,t).
Dividing these equations by Δχ and letting Δχ go to zero, we arrive at two differential equations:
(7.78)
∂e∂x(x,t)=-Ri(x,t)-L∂i(x,t)∂t
(7.79)
∂i∂x(x,t)=-Ge(x,t)-C∂e(x,t)∂t.
Finally, we can find an equation for e ( x , t ) only by differentiating (Equations 7.78) and (7.79) with
respect to x and t, respectively, to eliminate i ( x , t ) . We obtain
(7.80)
exx=LCett+(LG+RC)et+RGe.
Similarly, we can show that i ( x , t ) satisfies
(7.81)
ixx=LCitt+(LG+RC)it+RGi.
(Equations 7.80) and (7.81) are known as the telegraph equations.
Special Cases
High‐Frequency Limit. To qualitatively analyze this limit, consider the case where
(7.82)
e ( x , t ) = A ( x ) sin ( ω t + ϕ 1 ) , i ( x , t ) = B ( x ) sin ( ω t + ϕ 2 )
and ω ≫ 1. Under these assumptions, the second term in the right‐hand side of (Equation 7.78),
whose ((effective coefficient” is Lω (the impedance), is much larger than the first term, whose
effective coefficient is R. Hence, (Equation 7.78) can be approximated in this limit by
(7.83)
ex=-Lit.
Similarly, (Equation 7.79) reduces to
(7.84)
ix=Cet.
Combining (Equations 7.83) and (7.84) we obtain the wave equation
(7.85)
exx=LCett,ixx=LCitt.
Low‐Frequency Limit. In this case, i and e change very slowly with time. There‐ fore, from a similar
qualitative analysis as performed for the high frequency limit, with ω ≪ 1, we obtain that the
effective coefficient of the second term (Lω) is much smaller than R. Consequently, we can
approximate (Equations 7.78) and (7.79) by
(7.86)
ex=-Ri,ix=-Ge
and therefore,
(7.87)
exx=RGe,ixx=RGi.
These are ordinary differential equations for i and e.
Submarine Cable. In the 19th century, telecommunication signals between the United States and
Europe were transmitted by cables that were laid down on the ocean floor. For these cables, G ≅ O
because of their extreme insulation. Moreover, the signal frequency ω is low. Under these
circumstances, we infer from (Equation 7.82) that the impedance Lω is much smaller than R. Hence,
(Equation 7.78) can be approximated by
(7.88)
ex=-Ri.
Furthermore, since we have assumed G = 0, (Equation 7.79) simplifies to
(7.89)
ix=-C∂e∂t.
Combining these two equations, we can approximate (Equations 7.78) and (7.79) by
(7.90)
exx=RCet,ixx=RCit
which show that e and i satisfy the one‐dimensional diffusion equation.
Exercises
1. Give explicit derivations of (Equations 7.80) and (7.81).
2. Compare (Equations 7.80) and (7.81) with the model equation for a spring mass system with
friction and identify the physical meaning of each term in these equations.
3. Explain in detail the approximations that lead to (Equations 7.83), (7.84), (7.86), (7.89),
and (7.90).
4. Derive a differential equation for the voltage e ( t ) in the circuit in Fig. 7.13 if the current i
( t ) is known.
Figure 7.13 LRC circuit
As a corollary, we observe that if Φ 1 and Φ 2 are the potential functions for the gravitational field
of the masses M 1 and M 2, respectively, then
(7.101)
Ftotal=-∇ Φ1-∇ Φ2=-∇ (Φ1+Φ2)
where
(7.102)
Φi=-GMirii=1,2.
Thus, the potential function of the total gravitational field is given by the (scalar) sum of the
individual potential functions.
Similarly, if we are given a finite number of point masses, M i , i = 1, n with gravitational fields F
i and potential functions Φ i , then
(7.103)
Ftotal=∑i=1nFi=-∑i=1n∇ Φi=-∇ (∑i=1nΦi).
Remark 7.5.3
If the total potential function Φ of a gravitational field is known, then the gravitational field itself
is given as F = V$. This is one of the reasons for the introduction of the potential function, since Φ
is simply a (scalar) sum of the individual potentials and hence easier to compute than the (vector)
sum of the gravitational fields.
Idealizations
1. We assume that the concept of a point mass is valid. As a matter of fact, we note that due to
the discrete nature of matter, the notion of a (mathematical) point particle with mass m has
no physical meaning.
2. We assume that the field generated by a point particle does not act on itself; otherwise,
various contradictions will creep in.
Modeling
To compute the gravitational field due to a continuous mass distribution with mass density ρ(x) in a
volume V, we divide the volume into small cells of volume ▵ V i . If we consider each of these cells
as a point particle of mass ρ ( x i ′ ) △ V i (where x i ′ ε △ V i ), then the gravitational field due to it
at a point
x = ( x , y , z ) is
(7.104)
△ Fi=-Gρ(xi′ )△ Viri2eri
where
(7.105)
ri=|x-xi′|=(x-xi′)2+(y-yi′)2+(z-zi′)2
(see Fig. 7.17).
Discussion
One possible approach to modeling the traffic flow is to describe each car as a finite element on
the highway and then write a discrete model, which describes the motion of each such car. However,
if there are many cars on the highway, this approach is not practical, and it is better to construct a
continuous model, which treats these cars as “smeared out” quantities. We construct such a
continuous model in this section.
Approximation and Idealization
We assume that the highway is infinite.
1. We define the car density p ( x,t ) as
(7.128)
ρ(x,t)≅ Numberofcarsontheinterva1[x,x+△ x]attimet△ x
where Δχ must be large compared to a car length. [Otherwise, p(x, t) = 1 if there is a car at x in time
t, or p(x,t) = 0 if there is none.] In fact, the same approximation is made whenever we define the mass
density of a “continuous” body made of discrete atoms and molecules. Hence, the equations we
derive in this section apply also to fluid flow in a pipe.
1. We assume that there are no accidents on the highway (or that their number is negligible).
Hence, we can formulate the principle of “car conservation” (which is equivalent to that of
mass conservation) as follows:
The rate at which the number of cars on the segment [a,b] is changing equals the rate at which
they enter less the rate at which they are leaving.
1. We define the concept of car flux q ( x, t ) in the same way that we defined this concept in
Section 2. However, in this context it is natural to define the flux per lane rather than per
unit area. Equivalently, this can be considered as letting the unit length be equal to the width
of the lane. Moreover, note that
(7.129)
q(x,t)=ρ(x,t)u(x,t)
where u ( x, t ) is the car’s speed at x at time t, and the dimension of q is
(7.130)
q=CarsTime.
Modeling
To derive a model equation for the traffic flow, consider a finite section of the road between a and
b . The number n of cars in this segment at time t is
(7.131)
n(t,a,b)=∫abρ(x,t)dx.
Hence, the rate of change in this quantity is
(7.132)
dndt=∫ab∂ρ(x,t)∂tdx.
This rate of change must equal the flux of cars entering at a less the flux of cars leaving at b .
(Remember that the flux was defined per lane!) Therefore,
(7.133)
dndt=q(a,t)-q(b,t)=-∫ab∂q∂x(x,t)dx.
Thus, we infer from (Equations 7.132) and (7.133) that
(7.134)
∫ab(∂ρ∂t+∂q∂x)dx=0.
(7.135)
∂ρ∂t+∂q∂x=0.
Using (Equation 7.129) to substitute for q ( x , t ) , we finally obtain
(7.136)
∂ρ∂t+∂(ρu)∂x=0.
This equation is the continuity equation in one dimension. Notice, how‐
ever, that this equation contains two unknown quantities, ρ and u. Therefore, to solve it we must
either be able to express u = u(ρ) or find an additional equation that relates these two quantities.
Compounding
To derive the version of the continuity equation in three dimensions, we consider a fluid flow with
mass density ρ ( x , t ) . Let V be a volume contained in the flow. The mass of the fluid in V at time t is
given by
(7.137)
m(t,V)=∫Vρ(x,t)dx.
Hence, the rate of change of mass in V is
(7.138)
dmdt=∫V∂ρ∂t(x,t)dx.
Now let S denote the boundary of V and n(x) the unit outward normal to S at x. The total mass flow
rate of the fluid across S in the outward direction is
(7.139)
∫Sq·ndS=∫S(ρu·n)dS.
The mass conservation principle implies, however, that the rate of change of mass in V must equal the
rate at which the mass is crossing S in the inward direction. Therefore,
(7.140)
∫V∂ρ∂t(x,t)dV=-∫Sρu·ndS.
To convert the right‐hand side of (Equation 7.140) into a volume integral, we now invoke the
divergence theorem, which states that for any smooth vector field F in V
(7.141)
∫SF·ndS=∫VdivFdV.
This yields
(7.142)
∫V[∂ρ∂t+div(ρu)]dV=0.
Since V is arbitrary, we infer that the integrand in (Equation 7.142) must be zero, or
(7.143)
∂ρ∂t+div(ρu)=0.
(Equation 7.143), which is a first‐order partial differential equation, is called the continuity equation
in three dimensions.
Exercises
1. From your experience, guess the general form of the relationship between u and p in a one-
lane highway.
2. Derive a model equation for an infinite one-lane highway where cars are entering and
leaving the highway at constant rates a and β per mile, respectively. Generalize to the case
where a = a ( x,t ) and β = β(x,t).
3. Compound the model in Exercise 2 to include accidents at a rate a ( x, t ) per mile.
4. Consider fluid flow in a long cylindrical pipe with constant cross section A whose axis is
along the x-axis. Let p ( x,t ) be the density of the fluid and q ( x, t ) be its flux. If the walls
of the pipe are made of porous material that allows the fluid to leak out at a rate L per unit
length, show that
∂ρ∂t+∂q∂x+LA=0.
1. Derive model equations for the car densities p1(x,t) and p2(x,t) in a two-lane infinite
highway with no entries or exits where cars are moving from lane 1 to lane 2 at a rate of a (
p 2 ) per mile and at a rate of b(p1) per mile from lane 2 to lane 1.
2. Compound the model of Exercise 5 to include entries and exits.
7.7 ELECTROMAGNETISM
To close these equations we need relations, D = D(E), H = H(B), and J = J(E) . In vacuum or in
homogeneous isotropic medium, we have the linear relations
(7.148)
D=εE,H=Bμ
where ɛ and μ are called the permittivity and permeability of the medium. Furthermore, in a
homogeneous conducting medium
(7.149)
J=σE
where σ is the medium conductivity. In the following, we assume the relations, (Equations 10.93) and
(10.94) implicitly.
By taking the divergence of (Equation 10.92) and using (Equation 10.89), we obtain the (electric
charge) continuity equation
(7.150)
∂ρ∂t+∇ ·J=0.
Remark 7.7.1
If a point charge ρ is moving with velocity y, then the induced current density is J = ρv.
Remark 7.7.2
The total force exerted by an electromagnetic field on charges and currents in a volume V is given
by the Lorentz force equation
(7.151)
F=∫V(ρE+J×B)dV.
Remark 7.7.3
To treat the special electromagnetic fields discussed in the rest of this section, we need the
following facts:
∇ × (∇ × A) = ∇(∇ ⋅ A) - ∇2 A
If ∇ × A = 0, then A is conservative and therefore there exists a scalar function ϕ so that
(7.153)
A=∇ ϕ.
If ∇ ⋅ A = 0, then A is solenoidal and, therefore, there exists a vector potential B so that
(7.154)
A=∇ ×B.
Furthermore, B can be chosen so that ∇ ⋅ B = 0
7.7.4 Magnetostatic
The equations for the magnetic field in this case are
(7.166)
∇ ·B=0
(7.167)
∇ ×B=μJ.
From (Equation 7.166), we infer that B is a solenoidal vector field. Hence, there exists a vector
potential A so that
(7.168)
B=∇ ×A.
Hence, from Equation (7. 167)
(7.169)
∇ ×(∇ ×A)=μJ.
If we choose A so that ∇ ⋅ A = 0, we obtain
(7.170)
∇ 2A=-μJ,
i.e., A satisfies a vector Poisson equation. Hence,
(7.171)
A=μ4π∫VJ(r′ )dr′ |r-r|.
By taking the curl of this equation, we obtain Biot‐Savart law
(7.172)
B=μ4π∫VJ(r′ )×(r-r′ )|r-r|3dr.
Since the total magnetic charge is always zero, the multipole expansion for B yields
(7.173)
B=μM4πr5
where
(7.174)
Mi=∑j(3rirj-r2δij)mj
(7.175)
m=12∫Vr′ ×J(r′ )dr′ .
CONTENTS
8.1 Method of Separation of Variables
8.1.1 Method of Separation of Variables By Example
8.1.2 Non Cartesian Coordinate Systems
8.1.3 Boundary Value Problems with General Initial Conditions
8.1.4 Boundary Value Problems with Inhomogeneous Equations
8.2 Green’S Functions
8.3 Laplace Transform
8.3.1 Basic Properties of the Laplace Transform
8.3.2 Applications to the Heat Equation
8.4 Numerical Solutions Of PDES
8.4.1 Finite Difference Schemes
8.4.2 Numerical Solutions for the Poisson Equation
8.4.2.1Other Boundary Conditions
8.4.3 Irregular Regions
8.4.4 Numerical Solutions for the Heat and Wave Equations
8.1 Method of Separation of Variables
When it comes to solving boundary value problems involving partial differential equations, a number
of approaches are available. The method of separation of variables is very convenient because it
draws on many well‐known mathematical concepts and frequently works well. In this section, we
introduce this method and illustrate its application by considering various systems that were modeled
in the last chapter by partial differential equations. In particular, we concentrate on the heat, wave,
and potential equations that are important in many scientific and engineering applications.
The general objective of this method is to reduce the solution of a given partial differential
equation into the solution of a number of ordinary differential equations. Very often these ordinary
differential equations are well known, and their solutions are easily found. The whole process
follows a logical step‐by‐step development that terminates in the evaluation of the Fourier
coefficients of a Fourier‐type series. We suggest that you carefully catalog each step and its position
in the stairway to a successful conclusion.
Solution 8.1.3 In order to carry out specific operations, let us examine the special case where f
1(x) = 0 and f 2(x) = x. Steps 1,2, and 3. Assume ( x , y ) = X ( x ) Y ( y ) . Computing the necessary
derivatives and substituting into the diffe rential equation, we have
X′ ′ Y+XY′ ′ =0
or
X′ ′ X=-Y′ ′ Y=-λ
which yields the two ordinary differential equations
X′ ′ (x)+λX(x)=0
Y′ ′ (y)-λY(y)=0.
Steps 4 and 5. From our assumption and the given boundary conditions, we arrive at the following
boundary conditions associated with the ordinary differential equations:
.X(0)=0, X(a)=0
and
Y(0)=0.
Case 1. λ > 0 . Let λ = α 2 > 0 . Then we can show as before that the eigenvalues are
λn=αn2=n2π2a2n=1,
and the eigenfunctions are
X n ( x ) = sin n π x a .
Cases 2 and 3. λ = 0 and λ < 0 . It is easily shown that there are no non‐positive eigenvalues.
Step 6. One new idea introduced in this problem occurs when we wish to solve the second
differential equation,
Y′ ′ -n2π2a2Y=0.
Normally a beginning student would solve this equation using exponentials as follows
(8.13)
Yn(y)=Cnexpnπya+Dnexp(-nπya)
where e x p y = e y . Because neither exponential in Equation (12.37) vanishes, it is necessary to
solve a system of two equations to find D in terms of C. Therefore, it is much more convenient to
use hyperbolic functions, and we can write
Y n ( y ) = C n cosh n π y a + D n sinh n π y a .
Using the boundary condition Y(0) = 0, we see that
Yn(0)=0=Cn·1+Dn·0
or
Cn=0
and that
Y n ( y ) = sinh n π y a
is a solution to our ordinary differential equation in y and its one boundary condition. For
convenience we have again set the constant D = 1.
Step 7. Combining the two families of solutions X(x) and Y n (y), we construct solutions to the
partial differential equation and the three homogeneous boundary conditions that take the form
u n ( x , y ) = X n ( x ) Y n ( y ) = sin n π x a sinh n π y a .
Step 8. We still have to meet the final inhomogeneous boundary condition u ( x , b ) = x . In order
to do this we use the superposition principle and form the linear combination
(8.14)
u ( x , y ) = ∑ n = 1 ∞ b n sin n π x a sinh n π y a
expecting that we can find values for b n such that the inhomogeneous condition is met. Letting
y = b in Equation (9.69), we see that
(8.15)
u ( x , b ) = x = ∑ n = 1 ∞ b n sinh n π b a sin n π x a
If we let
B n = b n sinh n π b a
Equation (9. 70) becomes
x = ∑ n = 1 ∞ B n sin n π x a
which is a Fourier sine series. Therefore,
(8.16)
B n = b n sinh n π b a = 2 a ∫ 0 a x sin n π x a d x
= 2 a a 2 n 2 π 2 [ sin n π x a - n π x a cos n π x a ] | 0 a
= 2 a n 2 π 2 [ - n π cos n π ] = ( - 1 ) n + 1 2 a n π .
The constants b n become
b n = ( - 1 ) n + 1 2 a n π sinh n π b a
and the final answer can be written as
u ( x , y ) = 2 a π ∑ n = 1 ∞ ( - 1 ) n + 1 sin ( n π x / a ) sin h ( n π y / a ) n sin h ( n π b / a ) .
Remark.
When solving the ordinary differential equation
y′′(x)-α2y(x)=0
it is useful to remember that there are three practical ways in which to express the solution. If one of
the boundary conditions is of the form
y(0) = 0 or y ’(0) = 0
then the general solution
(8.17)
y ( x ) = A cosh α x + B sinh α x
is most convenient to use because the value of A or B is quickly determined. On the other hand, if
y(L) = 0 or y ’(L) = 0
is given as a boundary condition, one of the following forms of the solution is useful:
y ( x ) = E sinh α ( x - d ) f o r y ( L ) = 0
y ( x ) = E cosh α ( x - d ) f o r y ( L ) = 0
where E and d are arbitrary constants. These forms of the solutions are not as well known as the form
in (Equation 9.71), but they can easily be shown to be equivalent for the particular boundary
condition given using the well‐known hyperbolic identities
sinh ( α - β ) = sinh α cosh β - cosh α sinh β
and
cosh ( α - β ) = cosh α cosh β - sinh α sinh β .
Finally, the classic solution
y(x)=Aeαx+Be-αx
finds its greatest use in solving boundary value problems over semi‐infinite intervals and the study of
Fourier integrals.
Exercises
1. The temperature u ( x , t ) in a laterally insulated rod of length L satisfies the following boundary
value problem:
(8.18)
∂2u∂x2=1k∂u∂t0<x<L,0<t
u(0,t)=00<t
u(L,t)=00<t
u ( x , 0 ) = 100 0 < x < L .
Use the technique of separation of variables to find u ( x , t ) .
2. Use the technique of separation of variables to solve for the temperature u ( x , t ) if
(8.19)
∂ 2 u ∂ u 2 = 1 k ∂ u ∂ t 0 < x < 10 , 0 < t
∂u∂x(0,t)=00<t
∂ u ∂ x ( 10 , t ) = 0 0 < t
u ( x , 0 ) = 1 - x 0 < x < 10 .
3. The voltage e ( x , t ) along a submarine cable 3000 kilometers long satisfies the boundary value
problem
(8.20)
e x x = R C e t 0 < x < 3000 , 0 < t
e(0,t)=00<t
e ( 3000 , t ) = 0 0 < t
e ( x , 0 ) = sin x 100 0 < x < 3000 .
Use the technique of separation of variables to find e ( x , t ) .
4. The current i ( x , t ) along a submarine cable of length L satisfies
(8.21)
ixx=RCit0<x<L,0<t
∂i∂x(0,t)=00<t
∂i∂x(L,t)=00<t
i(x,0)=20<x<L.
Use the technique of separation of variables to find the current i ( x , t ) in the cable.
5. Find the temperature u ( x , t ) by the technique of separation of variables in a laterally insulated
rod of length 20 meters that has a heat source given by δ u ( x , t ) j o u l e s / m 3 . The other
conditions are
(8.22)
u(0,t)=0
(8.23)
u ( 20 , t ) = 0
(8.24)
u ( x , 0 ) = x ( 20 - x )
6. The voltage e ( x , t ) satisfies the differential equation e xx = RCe t + Ae x . Using separation of
variables find e ( x , t ) if
(8.25)
∂e∂x(0,t)=00<t
∂e∂x(L,t)=00<t
e ( x , 0 ) = 50 0 < x < L .
7. (a) A vibrating string is fastened to air bearings situated on vertical rods at x = 0 and x = 2. Find
the displacement y ( x , t ) if the conditions are
(8.26)
yx(0,t)=00<t
yx(2,t)=00<t
y(x,0)=x0<x<2
∂y∂t(x,0)=00<x<2
(b) Sketch y ( x , t ) over 0 < x < 2 for t = 0, t = 1, and t = 2.
Hint. Use only the first two terms of the series solution.
8. The pressure p ( x , t ) in an organ pipe satisfies the differential equation
pxx=1c2ptt.
If the pipe is L meters long and open at both ends, find the pressure p ( x , t ) if p ( x , t ) = 0 and ( ∂ p
/ ∂ t ) ( x , 0 ) = 40 .
9. (a) Given the telegraph equation for finding voltage e ( x , t ) on a transmission line of length a
along with the boundary and initial conditions, we can write
(8.27)
exx=LCett+(RC+GL)et+RGei
0<x<a,0<t
∂e∂x(0,t)=0,0<t
∂e∂x(a,t)=0,
e ( x , 0 ) = 100 , 0 < x < a
∂e∂t(x,0)=00<x<a.
Solve for e ( x , t ) using separation of variables.
(b) What is the frequency of the third harmonic?
10. The length of a guitar string is 65 centimeters. If the string is plucked 15 centimeters from the
bridge (i.e., at the end of the wire) by raising it 3 millimeters, find the displacement u ( x , t ) using
separation of variables.
11. The length of a piano string is 1 meter. When a pupil strikes a key, the following velocity is
imparted to the string:
∂ u ∂ t = velocity
(8.28)
= 0 0 < x < 49 c m 1 49 < x < 51 c m 0 51 < x < 100 c m .
Find the displacement u ( x , t ) using separation of variables. Do not evaluate the Fourier coefficients
of the final solution.
8.1.2 Non Cartesian Coordinate Systems
Examples 1, 3, and 4 were chosen especially so that you would see the method of separation of
variables applied to the three classes of second‐order differential equations: parabolic, hyperbolic,
and elliptic. All were set in terms of rectangular coordinates. In Example 5 we will examine a
boundary value problem that is more easily solved in polar coordinates.
Example 8.1.4 : Laplace Equation on a Disk
Let us consider a problem similar to Example 4 except that the shape of the plate is circular
rather than rectangular. We wish to find the steady‐state temperature u ( r , θ ) throughout a
circular plate of radius c that is insulated laterally. The temperature on the circumfe rence is 1000
C over one semicircle and 0 o over the other (see Figure 8.2).
Solution 8.1.5 To solve this problem we introduce polar coordinates and note that in these
coordinates the velocity is given by
(8.41)
v=vφ=φrer+1rφθeθ
where e r is the unit vector outward from the origin and e θ , a unit vector, is perpendicular to e r in
a counterclockwise direction.
Looking at Figure 8.4 we observe that the boundary conditions at r = + ∞
Figure 8.4 Radial coordinate system and the components of φ
are
(8.42)
φ r = a cos θ , 1 r φ θ = - a sin θ
Since the fluid cannot penetrate the cylinder, we must have φ r ( c , θ ) = 0 . Finally, since the
velocity at θ and θ + 2π is the same, we must have from Equation (8.41)
(8.43)
φr(r,θ+2π)=φr(r,θ)
φθ(r,θ+2π)=φθ(r,θ)
The method used in this problem is similar to that used in Example 5. We assume φ ( r , θ ) = R ( r )
Θ ( θ ) , which (aft er substituting into Laplace’s equation in polar form) yields the two ordinary
diffe rential equations and boundary conditions
(8.44)
Θ′ ′ +λΘ=0,Θ(θ+2π)=Θ(θ),Θ′ (θ+2π)=Θ′ (θ)
(8.45)
r2R′ ′ +R′ -λR=0,R′ (c)=0
If λ = α 2 > 0 , then the general solution of Θ ’’ + α 2 Θ = 0 is
Θ ( θ ) = A cos α θ + β sin α θ
and
Θ ′ ( θ ) = - α A sin α θ + α B cos α θ .
Substituting these two equations into the periodic conditions in Equation ( 8.42 ), we find (aft er
some effort) that
α = n, n = 1, 2, 3, . . .
or
λn=n2
and
Θ n ( θ ) = A n cos n θ + B n sin n θ .
With these eigenvalues the diffe rential equation in Equation ( 8.45 ) becomes
r 2 R ’’ + rR ’ - n 2 R = 0n = 1, 2, . . .
whose solution is
Rn(r)=Cnrn+Dnr-n
Using the boundary conditions in Equation ( 8.45 ), we have
R′ (c)=0=nCncn-1-nDnc-n-1
from which it follows that
Dn=c2nCn.
For λ = 0 , Θ ’’ = 0, and its solution is
Θ0=A0+B0θ.
The solution to the diffe rential equation in Equation ( 8.45 ) is
Now since ln r → + ∞ as r → + ∞, we must set C 0 = 0 and we can let
R0=D0=1.
Since there are no negative eigenvalues, we expect the solution to take the form
(8.46)
φ ( r , θ ) = A 0 + B 0 θ + ∑ n = 1 ∞ [ r n + c 2 n r - n ] [ A n cos n θ + B n sin n θ ]
We still have two conditions [Equation ( 8.42 ] that must be met. Differentiating Equation ( 8.46 )
with respect to r and θ, we have
(8.47)
φ r = ∑ n = 1 ∞ [ n r n - 1 - n c 2 n r - n - 1 ] [ A n cos n θ + B n sin n θ ]
and
(8.48)
φ θ = B 0 + ∑ n = 1 ∞ [ r n + c 2 n r - n ] n [ - A n sin n θ + B n cos n θ ]
Now the only term in Equation ( 8.47 ) that is bounded occurs when n = 1; therefO re, A n , B n = 0
for n = 2, 3, . . . and
lim r → ∞ φ r ( r , θ ) = a cos θ = A 1 cos θ + B 1 sin θ
or
A1=a
B1=0.
Under these conditions Equation ( 8.48 ) becomes
(8.49)
1 r φ θ = B 0 r + 1 r [ r + c 2 r - 1 ] [ - A 1 sin θ ]
= B 0 r + [ 1 + c 2 r 2 ] [ - a sin θ ] .
As r → + ∞,
1 r φ θ = - a sin θ
and the second condition in Equation ( 8.42 ) is satisfied.
Combining all these conditions in Equation ( 8.46 ), we write the solution as
φ ( r , θ ) = A 0 + B 0 θ + a [ r + c 2 r ] cos θ
where A 0 and B 0 are arbitrary constants.
As we investigate this solution, we must recall that φ ( r , θ ) is the velocity potential, not the
velocity. The velocity field is given by
y = φ r e r + 1 r φ θ e θ = a [ 1 - c 2 r 2 ] cos θ e r - a [ 1 + c 2 r 2 ] sin θ e θ .
When we measure the velocity far from the axis of the cylinder,
v ≈ a cos θ e r —asin θe θ = ai.
In other words, the effect on the velocity due to the cylinder becomes less noticeable as we move
away from the axis. On the other hand, when we are near the cylindrical obstruction, that is r ≈ c,
v ≈ - 2 a sin θ e θ
which shows that the flow follows the shape of the cylinder when r ≈ c and approaches zero as we
get nearer to the x‐axis.
Example 8.1.6 :Vibrations in a Sphere
In all our examples up to this point it has been relatively easy to evaluate the eigenvalues and
eigenfunctions. We will now examine a problem in which this is not the case. We wish to study the
vibrations or pressure of air within a sphere of radius c which might be caused by an exploding
firecracker, for example.
The pressure equation in space is given by
(8.50)
( ρ 2 p ρ ) ρ + 1 sin 2 θ p φ φ + 1 sin θ ( p θ sin θ ) θ = ρ 2 a 2 p t t
Solution 8.1.6 If we assume the pressure is independent of θ and ϕ(i.e., in the radial direction
only), Equation ( 8.50 ) becomes
(8.51)
1ρ2(ρ2pρ)ρ=1a2ptt
Since the pressure under consideration is within the sphere, the directional derivative of p in the
direction ρ must be zero on the boundary, or
pρ(c,t)=0t>0.
Choosing a trial solution of the form p ( ρ , t ) = R ( ρ ) T ( t ) , we arrive at the two ordinary diffe
rential equations using the method of separation of variables on Equation ( 8.51 ):
(8.52)
ρ2R′ ′ +2ρR′ +λρ2R=0R′ (c)=0
(8.53)
T′′+a2λT=0
The general solution of Equation ( 8.52 ) for λ = α 2 > 0 is
R ( ρ ) = A cos α ρ ρ + B sin α ρ ρ .
This solution can be found by replacing R(ρ) by S(ρ)/ρ, which transfO rms the diffe rential
equation into one with constant coefficients (see Exercise 16). Since p ( ρ , t ) must be bounded in
the sphere, it is necessary that A = 0. trigonometric equation is difficult to solve in order to find
the eigenvalues. We can easily
see that there are an infinite number of eigenvalues by solving the equation graphically, as
shown in Figure 8.5.
Figure 8.5 Eigenvalues μ n are given by the intersections of y = μ with tanμ
The first nonzero root is 4.49, the second 7.73, and so on. Notice in particular that as
μ = αc → ∞, the intersection points approach (2n + 1)π/2, n = 1, 2, . . .. Ta ble 1 lists the first eight
nonzero solutions done numerically. Using these values for μ we see that the eigenvalues λ satisfy
λn=αn2=μn2c2n=1,2,...
and the eigenfunctions are
R n ( ρ ) = sin ( μ n ρ / c ) ρ .
To see if λ = 0 is an eigenvalue, we solve the diffe rential equation
R′ ′ +2ρR′ =0
whose general solution is
R(ρ)=Aρ+B.
Once again since R(ρ) must be bounded in the sphere, A = 0 and R(ρ) = B. Since R ’(ρ) = 0 for any
ρ, it certainly follows that R ’(c) = 0. Therefore, λ = 0 is an eigenvalue and its corresponding
eigenfunctions R 0(ρ) = 1. This ends our search for eigenvalues, for you can show there are no
negative eigenvalues.
We now proceed to the solution of Equation ( 8.53 ). For λ = μ n 2 / c 2 , n = 1, 2, . . ., its
solution is
T n ( t ) = C n cos a μ n t c + D n sin a μ n t c
and for λ = 0 ,
T0(t)=C0+D0t.
Combining the solutions R and T, we see that the answer to our boundary value problem takes the
form
(8.54)
p ( ρ , t ) = C 0 + D 0 t + ∑ n = 1 ∞ sin ( μ n ρ / c ) ρ
× ( C n cos a μ n t c + D n sin a μ n t c )
In order to solve for constants C n and D n , n = 0, 1, 2, . . ., we need two initial conditions:
p ( ρ , 0 ) = f ( ρ ) and ∂ p ( ρ , 0 ) ∂ t = g ( ρ ) 0 < ρ < c .
When we attempt to solve for the coefficients, the series is not the Fourier series. But the family
of functions does possess an orthogonal property, and we can solve for the coefficients as follows:
(8.55)
C0=3c3∫0cρ2f(ρ)dρ,D0=3c3∫0cρ2g(ρ)dρ
C n = 2 ( 1 + μ n 2 ) c μ n 2 ∫ 0 c ρ f ( ρ ) sin μ n ρ c d ρ
D n = 2 ( 1 + μ n 2 ) a μ n 3 ∫ 0 c ρ g ( ρ ) sin μ n ρ c d ρ .
Exercises
1. You wish to find the temperature u ( ρ , θ ) in a laterally insulated pieshaped region of radius c.
The temperature satisfies the differential equation
ρ2uρρ+ρuρ+uθθ=0
and the boundary conditions
u ( ρ , 0 ) = 00 < ρ < c
u ( ρ , π 6 ) = 00 < ρ < c
u(c,θ)=θ0<θ<π6.
Use the technique of separation of variables to find the temperature
u(ρ,θ)
2. Find the steady‐state temperature u ( ρ , θ ) in a circular plate insulated laterally of radius 10 if the
temperature on the circumference is 3 - θ.
3. Find the electrostatic potential Φ ( ρ , θ ) on the half disk plate with radius 25cm with the
boundary conditions on the potential Φ as shown in Figure 8.6.
4. Show that (Equation 8.52) can be transformed into one with constant coefficients by using the
substitution R(ρ) = S(ρ)/ρ.
5. Evaluate and graph the velocity y of the fluid around the cylinder discussed in the previous
section for a = 1, r = 2c, θ = 0, π/4, π/2 radians.
6. Apply the method of separation of variables to Laplace’s equation in Example 5 to find the two
ordinary differential equations in (Equations 9.72) and (9.73).
7. Show that there are no negative eigenvalues to be found in the boundary value problem in
(Equation 8.44).
8. Prove that the solutions of (Equation 8.50) satisfy
∫ 0 c ρ sin μ n ρ c d ρ = 0
where tan μ n = μ n .
9. Show that
(a) ∫ 0 c sin μ n ρ c sin μ m ρ c d ρ = 0 m ≠ n
(b) ∫ 0 c sin 2 μ n ρ c d ρ = c 2 ( μ m 2 1 + μ m 2 ) n = 1, 2, . . .
where tan μ n = μ n .
10. Using the series in (Equation 8.54) and the fact that p ( ρ , 0 ) = f ( ρ ) , justify C 0 in (Equation
8.55).
Hint. Multiply both sides of (Equation 8.53) by ρ 2 dρ and integrate from 0 to c.
11. Using the conditions in Exercise 22, justify C n in (Equation 8.55).
Hint. Multiply both sides of (Equation 8.54) by ρ sin ( μ m ρ / c ) d ρ and integrate from 0 to c.
12. The coefficients D n can be found easily from the equation for C n in (Equation 8.55). Using the
condition ∂ p ( ρ , 0 ) / ∂ t = g ( ρ ) , prove D n in (Equation 8.55).
13. Find p ( ρ , t ) if c = 1 and
pρ(1,t)=0t>0
p ( ρ , 0 ) = 00 < ρ < 1
pt(ρ,0)=ρ0<ρ<12012<ρ<1.
14. Write out the boundary value problem for finding the pressure p ( ρ , t ) in a spherical region if the
pressure is zero at 100 meters from the center and the initial pressure is zero, and the rate of change
of pressure with respect to t is h(p) .
15. Consider two concentric spheres of radius a and b, a < b, respectively. If the pressure on the
inner sphere is 1 and the rate of change of pressure with respect to ρ on the outer sphere is 0, and the
initial pressure is zero and the rate of change of pressure with respect to t is ρ, write the partial
differential equation and boundary and initial conditions satisfied.
But
u3,1=u2,1=u1,1=u4,2=u4,3=0
and
u1,4=u2,4=u3,4=1.
Hence,
u2,3-4u2,2+u3,2+u1,2=0
u2,2-4u2,3+u3,3+u1,3=-1
u3,3-4u3,2+u2,2=0
u3,2-4u3,3+u2,3=-1
u1,3-4u1,2+u2,2+u0,2=0
u1,2-4u1,3+u2,3+u0,3=-1.
To solve the system we need two additional equations. These equations are obtained from the
boundary condition at ( 1 3 , 0 ) and ( 2 3 , 0 ) using the approximation
(∂u∂y)ij=ui,j+1-ui,j-12h.
This yields
u2,2-u0,2=23,
u2,3-u0,3=23.
The system now consists of eight equations in eight unknowns.
u1=u2=u3=u4
and so on.
Exercises
1. Derive a finite difference approximation scheme for the Poisson equation in three dimensions.
2. For a boundary value problem with circular symmetry in R 2, it is natural to use polar rather than
Cartesian coordinates. Derive a finite difference approximation to ∇2 u = f in these coordinates.
Hint: use a grid with constant ▵ θ and ▵ r. Note also that
∇ 2u=∂2u∂r2+1r∂u∂r+1r2∂2u∂θ2
3. Solve
∇ 2 u = r cos 2 θ u ( 1 , θ ) = 0
on the unit disk using ▵ θ = π/6 and ▵ r = 0.2. Compare with the exact solution.
4. Solve
∇ 2u=x2-y2
on
D={(x,y)0≤x≤1,0≤y≤1}
with the boundary conditions
∂u∂x(0,y)=1,u(x,O)=1,u(1,y)=u(x,1)=0.
Use different step sizes and compare the solutions obtained.
5. Complete the solutions of Examples 3,4, and 5. Use several different h.
6. Derive a finite difference scheme to solve
∇ 2u+(x2+y2)∂u∂x=f(x)
7. Solve numerically the differential equations that appear in the previous exercise if
f(x)=x2+y2
and the boundary conditions are the same as in Example 3.
Variational Principles
CONTENTS
9.1 Extrema of Functions
9.2 Constraints and Lagrange Multipliers
9.3 Calculus of Variations
9.3.1 Natural Boundary Conditions
9.3.2 Variational Notation
9.4 Extensions
9.5 Applications
9.6 Variation with Constraints
9.7 Airplane Control, Minimum Flight Time
9.8 Applications In Elasticity
9.9 Rayleigh‐ritz Method
9.10 The Finite Element Method in 2-D
9.10.1Geometrical Triangulations
9.10.2Linear Interpolation in 2D
9.10.3Galerkin Formulation of FEM
9.11 Appendix
9.1 Extrema of Functions
It is well known from elementary calculus that the local extrema of a smooth function f = f(x) in one
variable coincides with the points x i at which f ’(x i ) = 0. Furthermore, an extremal point x i is a local
maximum or minimum if f ’’(x i ) < 0 or f ’’(x i ) > 0 respectively.
Similar criteria exist naturally for multivariable functions. Thus, for functions in two variables F =
F ( x , y ) we have the following:
Definition: Let F ( x , y ) be C 2( = twice differentiable with continuous derivatives) in some region
R. The Hessian of F at ( x , y ) ε R is defined as
(9.1)
H(x,y)=Fxx(x,y)Fxy(x,y)Fyx(x,y)Fyy(x,y)
Remark: This definition can be extended naturally to functions with n variables. Thus if F = F ( x ) =
F ( x 1 , … , x n ) , then
(9.2)
H(x)=|∂2F(x)∂xi∂xj|
Definition: a = ( a 1 , … , a n ) is said to be a critical point of F(x) if
(9.3)
∂F∂xi(a)=0;i=1
, . . ., n
We quote the following theorem without proof:
Theorem 9.1.1 Let F ( x , y ) be C 2 in a domain D and let ( a , b ) be a critical point of F in the
interior of D. If
1. H ( a , b ) > 0 and F x x ( a , b ) < 0 then F has a maximum at ( a , b ) .
2. H ( a , b ) > 0 and F x x ( a , b ) > 0 then F has a minimum at ( a , b ) .
3. H ( a , b ) < 0 then F has neither a maximum nor a minimum at ( a , b ) .
Example: Let F ( x , y ) = x 2 - y 2 Then F x = 2x, F y = - 2y, and hence at (0, 0)F x (0, 0) = F y
(0, 0) = 0 ( i . e . ( 0 , 0 ) is a critical point of F) . However, H(0, 0) = - 4 < 0 and, therefore, (0, 0) is
not a maximum or a minimum point of F. To see what it is, we observe that when we approach (0, 0)
along the x‐axis (y = 0), F ( x , 0 ) = x 2 and hence (0, 0)“looks like” a minimum. On the other hand,
if we approach (0, 0) along the y‐axis then F ( O , y ) = - y 2 and looks like a maximum. Thus, (0, 0)
is a saddle point.
(9.12)
2y-2λx-2λy=0
(9.13)
2x2+2xy+y2=1
Solving (Equation 9.11) for y yields
(9.14)
y=-x(-1+2λ)λ
Substituting this expression for y in (Equation 12.9) and solving for λ we obtain
λ=3±52.
Using this expression of λ in (Equation 9.14) and then substituting for y in (Equation 12.10) we obtain
the following solutions for ( x , y ) :
( x , y ) = ( ± 0.528 , ± 0.325 ) , ( x , y ) = ( ∓ 0.850 , ± 1.376 ) .
The first two pairs represent the points with minimum distance from the origin while the last two
pairs have a maximum distance from the origin.
Remark: It should be observed from this example that the method of Lagrange multipliers is
“easiest to apply” when F is at most a quadratic polynomial and g k are linear functions since in this
case the solution of the resulting system of equations for the extremum points is straightforward.
Exercises
1. Plot the ellipse 2x 2 + 2xy + y 2 = 1 and identify the points closest to the origin.
2. Find the point on the ellipse x 2 - 2xy + 2y 2 = 1 with maximum distance from the origin.
3. Find the dimensions of the right circular cylinder of fixed total surface area A (including top and
bottom) with maximum volume.
4. Find the relative extrema of
F(x,y,z)=xz+yz
which lie on the intersection of the surfaces
x2+y2=2,yz=2
(two constraints).
5. Show that
(a) Among all triangles with the same perimeter the equilateral triangle has the greatest area.
(b) Among all rectangles with the same perimeter the square encloses the greatest area.
(9.23)
I′ (0)=∫x1x2[∂f∂YdYdε+dfdYdY′ dε]|ε=0dx=0
But from (Equation 12.28)
dYdε=η,dY′ dε=η′
and Y ( x , 0 ) = y ( x ) , Y ′ ( x , 0 ) = y ′ ( x ) ; therefore, (Equation 10.35) reduces to
(9.24)
∫x1x2[∂f∂yη+∂f∂y·η′]dx=0
Integrating the last term in (Equation 10.36) by parts we obtain
(9.25)
[∂f∂y·η]|x1x2+∫x1x2[∂f∂y-ddx[∂f∂y]]η(x)dx=0
Figure 9.1 Optimal trajectory and another one close by
9.4 EXTENSIONS
When the function f in the variational integral is a function of more than one dependent variable, e.g.
f=f( x,y,x˙ ,y˙ ,t) ,
the extremum of the variational integral
(9.61)
I=∫t1t2f(x,y,x˙,y˙,t)dt
is achieved by the functions x(t),y(t) satisfying the following Euler‐Lagrange equations.
(9.62)
∂f∂x-ddt[∂f∂x˙]=0
(9.63)
∂f∂y-ddt[∂f∂y˙]=0
Similarly for the n‐dimensional case (i.e. n‐dependent and one independent variables)
(9.64)
I=∫t1t2f(x,x˙,t)dt,xεRn
the extremum functions have to satisfy
(9.65)
∂f∂xi-ddt[∂f∂x˙i]=0,i=1,...,n
Another extension of the Euler‐Lagrange equations is required when f is a function of more than one
independent variable but one dependent variable, i.e.
(9.66)
I=∫∫Df(x,y,w,wx,wy)dA,DεR2
To prove the analog of (Equation 12.32) in this case let
(9.67)
W(x,y)=w(x,y)+εη(x,y)
where w ( x , y ) is the extremal function we are looking for.
(9.68)
0=dIdεε=0=∫∫D[∂f∂wη+∂f∂wxηx+∂f∂wyηy]dA
By applying Green’s theorem in two dimensions (see Appendix) we obtain
(9.69)
∫∫D[∂f∂wxηx+∂f∂wyηy]dA=
-∫∫Dη[∂∂x[∂f∂wx]+∂∂y[∂f∂wy]]dA
+∫∂Dη[∂f∂wxdy-∂f∂wydx]
Assuming that w is specified on the boundary of the domain ∂ D we deduce that η|∂D = 0. Hence the
extremum function w ( x , y ) has to satisfy the following partial differential equation:
(9.70)
∂f∂w-∂∂x[∂f∂wx]-∂∂y[∂f∂wy]=0
Exercises
1. Use variations on (Equation 12.33) to derive (Equation 12.34).
2. Find the functions x(t),y(t),z(t) which minimize the integral
(9.71)
I=∫t1t2x˙2+y˙2+z˙2dt
and satisfy the boundary conditions
(9.72)
x(ti)=xiy(ti)=yiz(ti)=zii=1,2
3. On a Riemanian manifold the infinitesimal distance is given by
(9.73)
ds2=gijxdxidxj
(implicit simulation o n i , j ) . To find the equation of the curve x(t) of minimum length that connects
two points on the manifold ( = geodesic curve) it is enough to minimize the square of the distance, i.e.
minimize
(9.74)
I(x)=∫t1t2gij(x)dxidtdxjdtdt
Derive the explicit form of Euler‐Lagrange equation for x(t) .
9.5 APPLICATIONS
A key point in the application of variational principles to mechanical (and other) systems is
Hamilton’s principle.
Definition: If V:R n → R the gradient of V is defined as
gradV = ∇ V = [ ∂ V ∂ x 1 , ...., ∂ V ∂ x n ]
Definition: A system is called conservative if the forces acting on the system can be expressed as
the gradient of some scalar function V which is called the potential of the force.
Hamilton Principle: For a conservative system the actual motion (or trajectory) will minimize the
variational integral
(9.75)
I=∫t1t2(T-V)dt
where T is the kinetic energy of the system. (T - V) is also called the Lagrangian of the system.
Remark: The kinetic energy of a point particle of mass m is 1 2 m v 2 where y is the velocity of the
particle.
Example 1: Derive the equations of motion for a system of two masses and three springs as shown
in the diagram:
Figure 9.2 System of two masses and three springs
Solution: Let x, y be the displacement of the masses from equilibrium at time t. The kinetic energy
is then given by
(9.76)
T=12mx˙2+12my˙2
The potential energy of the springs in this position is given by
(9.77)
V=12kx2+12k(y-x)2+12.ky2
Therefore the variational integral for this system is
(9.78)
I=∫t1t2[12m(x˙2+y˙2)-k(x2-xy+y2)]dt
This is a variational integral with one independent and two dependent variables. Euler‐Lagrange
(Equations 12.31)‐(12.32) then lead to
(9.79)
mx¨+2kx-ky=0
(9.80)
my¨+2ky-kx=0
Example 2: Vibrating Membrane (Drum head)
We consider a “flexible membrane” which is stretched over some region D in the x - y plane and
bounded by a curve C. Denote the deviation of the membrane from the x - y plane by w ( x , y , t )
(where we assume that on the boundary w ( x , y , t ) | C = 0 ). The membrane is set in motion by an
initial disturbance from the equilibrium position, and we want to derive an equation of motion for w (
x,y,t) .
Solution: The kinetic energy of the membrane is obviously given by
(9.81)
T=12∫∫Dρ(x,y)wt2dxdy
To evaluate the expression for the potential energy we first note that for conservative systems
V(b) - V(a) is the amount of work needed to take the system from state a to state b.
For our system we can assume that the amount of work needed to take the membrane from its
equilibrium state (x - y plane) to another is proportional to the difference in the surface area (Hooke’s
law in two dimensions). Hence
(9.82)
V(w)=V(w)-V(w=0)=k∫∫D{1+wx2+wy2-1}dxdy.
For small deflections we can use the approximation
(9.83)
1+u≅ 1+12u,|u|<<1
to obtain
(9.84)
V(w)=k2∫∫D(wx2+wy2)dxdy
The variational integral which we have to consider is therefore
(9.85)
I=12∫t1t2{∫∫D[ρwt2-k(wx2+wy2)]dxdy}dt
Euler‐Lagrange equations for this case yield (one dependent variable and three independent ones)
(9.86)
k(wxx+wyy)-ρ(x,y)wtt=0
or equivalently
(9.87)
k∇ 2w=ρwtt
which is the wave equation in two dimensions.
Exercises
1. Modify the discussion above so that gravity is included.
2. Derive the equations of motion for the double pendulum.
(9.91)
x ˙ 2 = ℓ 1 cos θ d θ d t + ℓ 2 cos ϕ d ϕ d t
(9.92)
y ˙ 2 = - ℓ 1 sin θ d θ d t - ℓ 2 sin ϕ d ϕ d t
The expressions for the kinetic and potential energy are
(9.93)
T=m12(x˙12+y˙12)+m22(x˙22+y˙22)
(9.94)
V = m 1 g ℓ 1 ( 1 - cos θ ) + m 2 g ( ℓ 1 + ℓ 2 - ℓ 1 cos θ - ℓ 2 cos ϕ )
3. Derive the equations of motion of the system shown in Figure 9 where m 1 is constrained to move
only in the vertical direction.
Hint: x 2 = ℓ 2 sin θ
(9.95)
y 2 = y 1 + ℓ 2 cos θ
(9.96)
T=12m1y˙12+m22(x˙22+y˙22)
(9.97)
V = 1 2 k ( y - ℓ 0 ) 2 + m 1 g [ - y + ℓ 0 ] + m 2 g [ ℓ o + ℓ 2 - ( y 1 + ℓ 2 cos θ ) ]
where ℓ0 is the natural length of the spring.
4. Solve the system described by (Equations 12.41)–(12.42) with m 1 ≠ m 2 and k 1 ≠ k 2 ≠ k 3.
(9.130)
∂K∂u-ddt[∂K∂u˙]=-λ1(T-t)-2λ2u(t)=0
Hence
(9.131)
u(t)=-λ1(T-t)2λ2
(and therefore, as expected u(T) = 0).
Substituting (Equation 9.131) in (Equations 9.126), (9.127) and (10.115) leads to the following
three equations in the unknowns T, λ 1 , λ 2 .
(9.132)
T2(λ1T+3λ2g)=-6λ2h
(9.133)
λ 1 2 T 3 = 12 λ 2 2 c 2
(9.134)
λ1T(λ1T+4λ2g)=-4λ2
From Equation (9. 133) we infer
(9.135)
λ1λ2=-23cT3/2;
hence for the optimal flight time (Equation 9.132) yields the equation
(9.136)
3gT2+6h=23cT3/2
Finally, using (Equations 9.131) and (9.135), we obtain an explicit expression for the thrust as a
function of time:
(9.137)
u(t)=3Tc[1-tT]
Exercises
1. Explain in detail the inversion of the integral in (Equation 9.123).
2. Reconsider the minimum lift off time problem when air drag has to be taken into account; i.e. the
equation of motion for the plane is given by
x¨=-g-αx˙+u(t)
where α is a positive constant.
3. Reconsider the problem of this section when one wants to minimize the fuel consumption (e.g.,
in lift off from the moon surface).
9.8 APPLICATIONS IN ELASTICITy
In this section we discuss the modeling of transverse vibrations in elastic bars and thin plates using
variational principles.
A. Transverse vibrations in an elastic bar.
Consider an elastic bar of constant cross section, density ρ per unit length, and length L. To model
the transverse vibrations of such a bar we ignore the possible distortion of the cross sections and
assume that any small section of the bar moves as a rigid entity.
If u ( x , t ) is the displacement from equilibrium, the total kinetic energy of the bar is given by
(9.138)
T=ρ2∫0Lu˙2dx
The strain potential energy due to the vibrations is given by
(9.139)
J=∫∫Dz2dydz
By Hamilton’s principle the motion of the bar will be an extremum of the variational integral
(9.140)
I(u)=12∫t1t2∫0L(ρu˙2-EJuxx2)dxdt
with the boundary condition
(9.141)
u=ux=0atx=0,L
(clamped rod).
To derive the appropriate Euler‐Lagrange equation for I(u) we introduce
(9.142)
u=u+εη
in (Equation 10.91) where u is the (sought for) extremal solution. Hence
(9.143)
I′(0)=∫t1t2∫0L[∂f∂u˙η˙+∂f∂uxx]ηxxdxdt=0
where
(9.144)
f=12[ρu˙-EJuxx2]
Using integration by parts and the boundary conditions we obtain
(9.145)
∫t1t2∫0L∂f∂u˙η˙dxdt=-∫t1t2∫0L∂∂t[∂f∂u˙]ηdxdt
(9.146)
∫0L∂f∂uxxηxxdx=∫0L∂2∂x2[∂f∂uxx]ηdx
Hence
(9.147)
∂∂t[∂f∂u˙]-∂2∂x2[∂f∂uxx]=0
which yields
(9.148)
ρ∂2u∂t2+EJ∂4u∂x4=0
To solve this equation we need the boundary conditions at 0, L and the initial position and velocity of
the bar.
B Transverse vibrations of thin plate.
Consider the transverse vibrations of a uniform thin plate over a domain
R with the boundary conditions
(9.149)
u=∂u∂n|∂R=0
The kinetic energy of the plate is
(9.150)
T=12ρ∫Ru˙2dA
and the total strain potential energy is
(9.151)
V=12D∫∫R[(∇ 2u)2-2(1-σ)(uxxyyy-uxy2)]dA
Here D is the “flexural rigidity” of the plate and σ is the “Poisson ratio” of the material (σ is related
to the relationship between the strain‐stress tensors in the material of the plate).
The variational integral for this problem is
(9.152)
I(u)=
12∫t1t2∫∫R{ρu˙-D[(∇ 2u)2-2(1-σ)(uxxuyy-uxy2)]}dAdt=
12∫t1t1∫∫RfdAdt
Letting
(9.153)
u=u+εη
we obtain
(9.154)
I′ (0)=∫t1t2∫∫R[∂f∂u˙η˙+∂f∂uxxηxx+∂f∂uyyηyy+∂f∂uxyηxy]dAd
t=0
Using integration by parts, Green’s theorem and the boundary conditions yields:
(9.155)
∫t1t2∫∫R∂f∂u˙η˙dAdt=-∫t1t2∫∫Rη∂∂t[∂f∂u˙]dAdt
(9.156)
∫∫R∂f∂uxxηxx=∫∫Rη∂2∂x2[∂f∂uxx]dA+∫C[ηx∂f∂uxx-η∂∂x[∂f∂ux
x]]dy
etc.
Combining all these results together we finally obtain
(9.157)
ρ∂2u∂t2+D∇ 2(∇ 2u)=0
This equation is referred to as the biharmonic equation.
9.9 RAYLEIGH‐RITZ METHOD
Variational formulation of differential equations can be used to obtain approximate solutions for these
equations. The basic technique is due to Rayleigh‐Ritz, and it was the “precursor” to the current
Finite Element Methods that are used for the numerical solution of partial differential equations in
various applications.
We present this method through two examples:
Example 1: The bending of an elastic bar under uniform loading.
In the previous section we derived a variational principle for the vibrations of an elastic bar. If we
apply on such a bar a uniform static loading, p, the appropriate variational principle for the static
shape of this bar will be
(9.158)
I(u)=∫0L[12EJ(u′′)2-pu]dx
Euler‐Lagrange equations then yield
(9.159)
EJuxxxx-p=0
If the rod is clamped at x = 0, the corresponding boundary conditions are
(9.160)
u(0)=u′(0)=0
We can solve (Equation 9.159) directly (as an ordinary differential equation) or use the variational
principle given by (Equation 9.158) to obtain at least an approximation to the solution. To this end we
consider (as an example) the function space S = { 1 , x , x 2 , x 3 , x 4 } and attempt to find the best
approximation to the solution in this space; i.e., we seek to find the values of a, b, c, d, e which yield
the best approximation to the solution u in the form
(9.161)
u=ax4+bx3+cx2+dx+e
These values of a, b, c, d, e will minimize I(u) in the function space S subject to the boundary
conditions.
As a first step toward the solution we apply the boundary conditions (Equation 9.160) to (Equation
9.161). This yields d = e = 0, i.e.
(9.162)
u(x)=ax4+bx3+cx2
Substituting (Equation 9.162) in (Equation 9.158) we obtain after integration:
(9.163)
I(u)=-pL3[15aL2+14bL+13c]+
E J [ 72 5 a 2 L 5 + 18 a b L 4 + 1 3 ( 24 a c + 18 b 2 ) ) L 3 + 6 b c L 2 + 2 c 2 L ]
This expression attains its minimum at a point where
(9.164)
∂I∂a=∂I∂b=∂I∂c=0
This leads to a system of three linear equations for a, b, c whose solution is
(9.165)
a = p 24 E J , b = - p L 6 E J , c = p L 2 4 E J
Example 2: Solve
(9.166)
∇ 2u=c
on the square Ω = [ - a , a ] × [ - a , a ] subject to the Dirichlet boundary condition
(9.167)
u|∂Ω=0
Solution: The solution of (Equation 9.166) satisfies the variational principle
(9.168)
I(u)=∫Ω[12(ux2+uy2)+cu]dxdy
Let {φ i }i = 1…N be a set of functions which satisfy the boundary condition (9.167). We seek an
approximate solution of (Equations 9.166), (9.167) in the form;
(9.169)
u(x,y)=∑aiϕi(x,y)
which minimizes the functional in (Equation 9.168), i.e.
(9.170)
∂I(∑aiϕi(x,y))∂ak=0k=1,...,N
This yields (after substitution in (Equation 9.168))
(9.171)
∑iai∫Ω[(∂ϕi∂x)(∂ϕk∂x)+(∂ϕi∂y)(∂ϕk∂y)]dxdy+c∫Ωϕkdxdy=0.
This is a system of N linear equations for the coefficients a i whose solution yields (through
Equation(9.169)) an approximate solution to (Equations 9.166), (9.167).
1. 1 4 4 - 2 - 2 - 2 2 0 - 2 0 2 u 1 u 2 u 3 = F 1 1 F 2 1 F 3 1
2. 1 4 2 - 2 0 - 2 4 - 2 0 - 2 2 u 2 u 4 u 3 = F 2 2 F 4 2 F 3 2
3. 1 4 4 - 2 - 2 - 2 2 0 - 2 0 2 u 3 u 4 u 5 = F 3 3 F 4 3 F 5 3
4. 1 4 - 2 - 2 0 - 2 4 - 2 0 - 2 2 u 4 u 6 u 5 = F 4 4 F 6 4 F 3 4
Combining these equations yields:
144-2-2000-240-200-208-4-200-2-480-200-204-2000-2-24u1u2
u3u4u5u6=F11F21+F22F31+F32+F33F42+F43+F44F53+F54F64
9.11 Appendix
Green’s Theorems in Two and Three Dimensions
Green’s theorems are the analog of integration by parts” in higher dimensions. However, some
variations of these theorems are useful in other contexts.
Theorem (Green’s Theorem in two dimensions) Let D be a region in two dimensions and let ∂ D be
its boundary (traced in the positive direction), then for any two (smooth) functions P, Q (and proper
assumptions on D), ∫ ∫ D [ ∂ P ∂ x + ∂ Q ∂ y ] d A = ∫ ∂ D (Pdy—Qdx).
Corollary: (Integration by parts)
If P = Gη. A = Fη then we obtain
∫∫D[G∂η∂x+F∂η∂y]dA=-∫∫Dη[∂G∂x+∂F∂y]dA+∫∂Dη(Gdy-Fdx)
Corollary: (Second order integration by parts)
If we let Q = 0 and P = [ G ∂ η ∂ x - η ∂ G ∂ x ] in Green’s theorem we obtain
∫∫DG∂2η∂x2dA=∫∫Dη∂2G∂x2dA+∫∂D(G∂η∂x-η∂G∂x)dy.
Similar equation holds for differentiation with respect to y.
Theorem: (Green’s theorem in three dimensions)
∫ V F · g r a d η d V = - ∫ V η divFdV + ∫ ∂ V η F · n d S .
Note that the divergence theorem is obtained as a special case of this theorem with η = 1.
CHAPTER 10
CONTENTS
10.1 Strain and Stress
10.2 Equations of Motion for Ideal Fluid
10.2.1Continuity equation
10.2.2Eulers’ equations
10.3 Navierstokes Equations
10.4 Similarity and Reynolds’ Number
10.5 Different Formulations of Navierstokes Equations
10.6 Convection and Boussinesq Approximation
10.7 Complex Variables in 2‐D Hydrodynamics
10.8 Blasius Boundary Layer Equation
10.9 Introduction to Turbulence Modeling
10.9.1Incompressible Turbulent Flow
10.9.2Modeling Eddy Viscosity
10.9.3k - ɛModel
10.9.4The Turbulent Energy Spectrum
10.10Stability of Fluid Flow
10.11Astrophysical Applications
10.11.1Derivation of the Model Equations
10.11.2Steady State Model Equations
10.11.3Physical Meaning of the Functions H(ρ),S(ρ)
10.11.4Radial Solutions for the Steady State Model
10.12Appendix A-Gauss Theorem And Its Variants
10.13Appendix B ‐Poincare Inequality And Burger’S Equation
10.14Appendix C ‐Gronwell Inequality
10.15Appendix D‐The Spectrum
10.1 Strain and Stress
By definition the relative positions of the points in a rigid body cannot change over time. For elastic
bodies, on the other hand, these relative positions can change. Strain measures the relative
deformation of the points in an elastic body when these deformations are “small.”
To analyze this concept, consider two nearby points whose coordinates in the undeformed position
are x = ( x 1 , x 2 , x 3 ) and y = ( y 1 , y 2 , y 3 ) . Let these points be displaced now to x+ d(x),
y+ d(y) (that is, the deformation is a function of the position).
The distance between the points before and after the deformation respectively is
(10.1)
r02=∑i=13(yi-xi)2
(10.2)
r12=∑i=13(yi-xi+di(y)-di(x))2
Since we are assuming that ∥ x‐y|| < < 1 ||d|| < < 1 we can write to a first order approximation
(10.3)
di(y)-di(x)≅ ∑j=13∂di∂xj(x)(yj-xj)
Expanding Equation (10.2) using Equation (10.3) and neglecting higher order terms we obtain
r12=r02+2∑i,j=13∂di∂xj(x)(yi-xi)(yj-xj)
We now define the extension of the element ||y—x|| as
e = lim y → x r 1 - r 0 r 0 = lim y → x ( r 1 - r 0 r 0 ) ( r 1 + r 0 2 r 0 ) =
= 1 2 lim y → x r 1 2 - r 0 2 r 0 2 = ∑ i , j = 1 3 ∂ d i ∂ x j ( x ) cos θ i cos θ j .
where we used the fact that
lim y → x ( r 1 + r 0 2 r 0 ) = 1 ,
and defined
cos θ i = lim y → x y i - x i r 0
are the cosine direction of the element. We therefore showed that the extension e is a quadratic form
in the direction cosines which can be rewritten as
(10.4)
e = ∑ i , j e i j cos θ i cos θ j
where
(10.5)
eij=12(∂di∂xj+∂dj∂xi).
Since e ij = e ji , this quadratic form is symmetric and e ij are called the components of the strain [the
symmetric strain matrix (e ij ) forms a second rank tensor].
Strains in elastic bodies are the result of forces acting on the boundary of each volume element.
The force per unit area is called the stress P. Observe however that in general P is not normal to the
surface element. The normal component of the stress is called the pressure (or tension), while the
tangential component is called the ((shearing stress To analyze the stress P acting on a surface
element dS with normal n we consider a tetrahedron of which three faces dS x , dS y , dS z lie in the
coordinate plane (see Figure 10.1).
Let τ i , i = 1, 2, 3, be the stresses on each face of the tetrahedron. By Newton’s third law the
condition of equilibrium for the volume enclosed by the tetrahedron is
(10.6)
PdS-τ1dSx-τ2dSy-τ3dSz=0
But
dS=ndS=dSxi+dSyj+dSzk,
i.e.
dSx=n1dS,dSy=n2dS,dSz=n3dS.
Therefore
(P-τ1n1-τ2n2-τ3n3)dS=0.
In component form this becomes
P 1 = τ 11 n 1 + τ 21 n 2 + τ 31 n 3
, etc.
Hence we can rewrite (Equation 10.6) as
(10.7)
Pi=∑j=13τjinj
It can be shown that (τ ij ) is a symmetric ( i . e . τ i j = τ j i ) second rank tensor which is called the
“stress tensor.”
Using the symmetry of (τ ij ) (Equation 10.7) can be written as
(10.8)
Pi=∑j=13τijnj
The basic assumption of linear elasticity is that for small deformations the strain and stress tensors
are linearly related. Thus
(10.9)
τij=∑m,nαijmnemn
where α ijmn depends on the elastic material.
Since e mn , τ ij are symmetric, it is easy to see that
(10.10)
αijmn=αijnm,αijmn=αjimn
Furthermore, it can be shown (using a thermodynamic argument) that
αijmn=αmnij
These constraints reduce the number of independent components in this tensor to 21. When the
material is isotropic (that is there is no preferred direction) the most general form of α ijmn is
(10.11)
αijmn=λδijδmn+μ(δimδjn+δinδjm)
where λ is the Lame’s constant and μ is the rigidity. Using (Equations 10.5), (10. 11) (Equation 10.9)
now becomes:
(10.12)
τij=λδij∑k∂dk∂xk+μ(∂di∂xj+∂dj∂xi)
10.9.3 k - ɛModel
As we saw in the previous section, zero and one equation models require the use of an empirical
“length scale.” As this quantity depends on the geometry and the boundary conditions, it is apparent
that a second equation either for this quantity or its “equivalent” is needed for a complete
specification of ν T . Two equation models, in spite of their deficiencies, are utilized today for most
turbulence research and applications. They can be used to compute properties of turbulent flow with
no apriori assumptions on the structure of the flow.
The most prominent among these two equation models is the k - ɛ model and its variants. Here k, ɛ
represent respectively the turbulent kinetic energy and dissipation per unit mass.
(10.144)
k=12[(u1′)2+(u2′)2+(u31)3]
(10.145)
ε=νT∂ui′∂xk∂ui′∂xk¯
(summation over i , k ) and ν T is given by
(10.146)
ν T = μ k 2 / ε , c = 0.09
The derivation of the equations for the evolution of these quantities is rather involved algebraically
and requires the modeling of terms containing various double and triple averages of ∂ u i ′ ∂ x j , e.g.
∂ui′∂xk∂ui′∂xm∂uk′∂xm¯.
The resulting evolution equations contain several constants which have to be determined
experimentally (and adjusted for different applications). Overall, the resulting model should be
considered “phenomenological.” Yet the model has been successful in many applications and is
currently the “industry standard” (we do not present this derivation here).
We should mention at this juncture that there exists an ongoing research effort to model the
evolution of turbulent flow using the Reynold stress tensor (Equation 10.140). This approach lessens
to some extent the need for the various modeling approximations which are made in the k - ɛ model.
However, in this approach six coupled partial differential equations have to be solved in addition to
the Navier‐Stokes equations. Such a scheme requires heavy computational efforts. Some attempts for
reductions based on some algebraic relations were suggested in the literature.
(10.162)
pi′ =χ(y)ei(αx+βz-ωt),α,β∈ R,ω∈ C
By a proper transformation in the x - z plane we can set β = 0 (i.e. let the x-axis coincide with the
wave front). Substituting in (Equations 10.154), (10. 155) we obtain
(10.163)
iαϕ1+dϕ2dy=0
(10.164)
-iωϕ1+iUαϕ1+ϕ2dUdy=-iαχ+1Re(d2dy2-α2)ϕ1
(10.165)
-iωϕ2+iαUϕ2=-dχdy+1Re(d2dy2-α2)ϕ2
(10.166)
-iωϕ3+iαUϕ3=-1Re(d2dy2-α2)ϕ3
(Equation 10.163) can be satisfied by introducing φ so that
ϕ2=iαϕ,ϕ1=dϕdy.
Also note that (Equation 10.166) is an “independent equation.”
Substituting φ in Equations(10.164), (10.165) we obtain
(10.167)
iϕ′ (-ω+αU)-iαϕdUdy=-iαχ+1Re(d2dy2-α2)ϕ′
(10.168)
-αϕ(-ω+U)=-χ′ +1Re(d2dy2-α2)(-iαϕ)
Differentiating (Equation 10.167) with respect to y and substituting in(Equation 10.168) for χ ’ we
obtain
(10.169)
i(U-c)(ϕ′ ′ -α2ϕ)-ϕU′ ′ =1αRe(ϕ(4)-2α2ϕ′ ′ +α2ϕ)
(where c = ω/α is the “phase velocity”).
(Equation 10.169) is called “Orr-Somerfeld equation.” It is a 4th order differential equation for φ
with the boundary conditions
ϕ(-h)=ϕ(h)=ϕ′(h)=ϕ′(-h)=0.
The stability of the solution to Tollmien-Schlichting wave is controlled by the
value of ω. When Imω > 0, the perturbation “explodes” and the solution is unstable. Neutral
stability is obtained when Imω = 0.
Im ω = ψ ( α , R e )
The curve ψ ( α , R e ) = 0 in the α - Re plane separates the domain of stability from the unstable
domain. It follows that there is a critical value of the Reynold’s number where the solution U
becomes unstable. At this point waves
start to appear in the flow. THEOREM (Rayleigh):
At Re = ∞, U(y) must have a point of inflection for flow instability.
Proof When Re = ∞, the right hand side Orr‐Somerfeld equation is zero, and we have
(10.170)
ϕ′ ′ -α2ϕ=U′ ′ U-cϕ
Taking the complex‐conjugate of this equation yields
(10.171)
ϕ′ ′ -α2ϕ∗ =U′ ′ U-c∗ ϕ∗
Multiplying (Equation 10.170) by φ * and (Equation 10.171) by φ and subtracting leads to
ddy(ϕ′ ϕ∗ -ϕϕ∗ ′ )=2iIm(c)U′ ′ |U-c|2|ϕ|2
Integrating this equation on [ - h , h ] and using the boundary conditions on φ we obtain
2iIm(c)∫-hhU′ ′ |ϕ|2|U-c|2dy=0.
Since all the terms in the integrand except U ’’ are positive, it follows that U ’’ must be 0 for some y ∈
[ - h,h] .
(10.182)
ρ(-ψxt-ψyψxx+ψxψxy)=-py-ρϕy+ρω2y
To eliminate p from these equations we differentiate (Equations 10.181), (10.182) with respect to y,
x respectively and subtract. This leads to
(10.183)
ρy(ψyt+ψyψyx-ψxψyy)+
ρ(ψyyt+ψyψyyx-ψxψyyy)-ρx(-ψxt-ψyψxx+ψxψxy)-ρ(-ψxxt-ψy
ψxxx+ψxψxxy)=-J{ϕ,ρ}+J{12ω2r2,ρ}
where r 2 = x 2 + y 2. The sum of the second and fourth terms in this equation can be rewritten as
(10.184)
ρ(∇ 2ψ)t+ρJ{∇ 2ψ,ψ}
To reduce the first and third terms in (Equation 10.183) we use (Equation 10.179). It follows that
(10.185)
ρy(ψyt+ψyψyx-ψxψyy)-
ρx(-ψxt-ψyψxx+ψxψxy)=ρyψyt+ρyψyψyx-(ρt+ρxψy)ψyy+ρxψx
t+(ψxρy-ρt)ψxx-ρxψxψxy=ρyψyt+ρxψxt-ρt∇ 2ψ+12J{(ψx)2+(
ψy)2,ρ}.
Combining the results of (Equations 10.184), (10.185), (Equation 10.183) becomes
(10.186)
ρyψyt+ρxψxt-ρt∇ 2ψ+ρ(∇ 2ψ)t+ρJ{∇ 2ψ,ψ}+12J{(ψx)2+(ψy)2
,ρ}=-J{ϕ,ρ}+J{12ω2r2,ρ}.
Thus we have reduced the original five (Equations 10.172)–(10.176) to three (Equations 10.176),
(10.179), and (10.186). Although (Equation 10.186) is rather cumbersome in general, it can be
simplified further when we consider only the steady state of the gas (a simplification for the time
dependent flow is also possible under some constraints but will not be presented here).
A strong dependence on ω is shown in Fig. 10.5 which has the same parameters as Fig. 10.4 except
that the boundary conditions on ρ are: ρ(0) = 0.35 and ρ(8) = 0.25. This figure clearly illustrates the
effect that rotation can have on the pattern of density fluctuations within the cloud. Furthermore, in this
figure the magnitude of the density fluctuations reverses itself as ω becomes larger viz. the higher
density peaks are placed at larger values of r (which is reminiscent of the situation in the solar
system).
Figure 10.5 Steady state of the interstellar gas with α = - 19.4, c = 1 and boundary conditions
ρ(0) =0.35, ρ(8) = 20.25 with different values of ω
10.12 Appendix A-Gauss Theorem And Its Variants
Notation: Let
x=(x1,x2,x3)=(x,y,z)
V = a three dimensional domain
∂ V = the boundary of V
n = ( n 1 , n 2 , n 3 ) the outward normal to V.
1. Basic Gauss Theorem: If φ(x) is a scalar field then
(A.1)
∫∂VϕnidS=∫V∂ϕ∂xidV
(this is basically the fundamental theorem of the calculus in 3-D).
2. Gradient form: Use the basic theorem for i = 1, 2, 3
∫∂Vϕn1dS=∫V∂ϕ∂x1dV
∫∂Vϕn2dS=∫∂ϕ∂x2dV
∫∂Vϕn3dS=∫∂ϕ∂x3dV.
Hence
(∫∂Vϕn1dS)i+(∫∂Vϕn2dS)j+(∫∂Vϕn3dS)k
=(∫V∂ϕ∂x1idV)+(∫V∂ϕ∂x2jdV)+(∫V∂ϕ∂x3kdV)
or in vector notation
(A.2)
∫∂VϕndS=∫VgradϕdV
3. Divergence form:
Let F = ( f 1 , f 2 , f 3 ) then from the basic theorem we have
∫∂Vf1n1dS=∫V∂f1∂x1dV
∫∂Vf2n2dS=∫V∂f2∂x2dV
∫∂Vf3n3dS=∫V∂f3∂x3dV.
Summing these three equations we have
(A.3)
∫∂VF·ndS=∫VdivFdV.
4. Tensor form:
Let T be a second rank tensor with vector components ( τ 1 , τ 2 , τ 3 ) then from (A.3) we have
∫∂Vτi·ndS=∫VdivτidV=
or in component form
∫∂Vτi·ndS=∫V∑j∂τij∂xjdV
or in tensor form
(A.4)
∫∂VT·ndS=∫VdivTdV.
5. Curl form:
∫∂Vf1n2dS=∫V∂f1∂x2dV
∫∂Vf2n1dS=∫V∂f2∂x1dV.
Hence
∫∂V(n1f2-n2f1)dS=∫V(∂f2∂x1-∂f1∂x2)dV.
This is the k component of
(A.5)
∫∂Vn×FdS=∫VcurlFdV.
CONTENTS
11.1 Atmospheric Structure
11.2 Thermodynamics and Compressibility
11.2.1 Thermodynamic Modeling
11.2.2 Compressibility
11.3 General Circulation
11.4 Climate
11.1 ATMOSPHERIC STRUCTURE
On the large scale (in height), the atmosphere is divided into three sections. These are
1. Homosphere (up to a height of 100km),
2. Heterosphere (100 to 500km),
3. Exosphere (above 500km)
In the exosphere, the air density is very low (and the mean free path is large). As a result,
molecules in this region have a “fair chance” to escape into space. In the heterosphere, the strong
ultraviolet radiation from the sun dissociates the H 2 O and O2 molecules. By this process, part of this
harmful radiation is filtered out and does not reach the lower levels.
In the homosphere, the molecular mean free path is small. As a result, bulk transport by turbulent
air motion dominates the diffusive processes. This turbulent mixing “homogenizes” the passive
constituents of the atmosphere;
i.e., their densities decrease exponentially with altitude at the same rate which gives air a
homogeneous composition of 78% N 2 and 21% O2.
Due to its importance, the homosphere is further divided into
1. Troposphere (between 0-10km in height),
2. Stratosphere (between 10-50km),
3. Mesosphere (above 50km).
Remark: The boundary layers between these are referred to as tropopause and stratopause. In the
troposphere (which is also referred to as the biosphere), the temperature decreases at a rate of 6.5°
Kelvin/km. In the stratosphere, on the other hand, the temperature increases with height due to ozone
heating (as a result, the stratosphere is “stably stratified; “ i.e. “lighter air” is on top of the “denser”
air). In the mesosphere, the temperature again decreases with height.
For a mole of an adiabatic parcel of gas (where no heat (Q) is exchanged between the parcel and
its surroundings), the first law of thermodynamics implies that
(11.3)
dQ=CVdT+PdV=0;
i.e. the sum of the change in the internal energy and the work done by the pressure is zero due to the
fact that there is no heat exchange with the environment. Using (Equation 12.1) leads to
(11.4)
CVdT+RTVdV=0.
Hence,
(11.5)
dTT=-RCVVdV.
Integrating this equation, we obtain
(11.6)
TVk-1=constant
where k = 1 + R C V . Using (Equation 12.1) we can rewrite (Equation 12.8) as
(11.7)
TP-k-1k=constant.
Taking the logarithmic derivative of (Equation 12.9) yields
(11.8)
dTT=k-1kdPP.
We assume now that the troposphere is composed of a gas which obeys the ((ideal gas law.” For an
adiabatic parcel of this gas betw TV^{{k - 1}} = een heights h and h + dh, the pressure difference
between the top and bottom is
(11.9)
dP=-ρgdh
where g is the acceleration due to gravity g = 9.8m/sec 2. Substituting for ρ from (Equation 12.2)
yields
(11.10)
dP=-gMPRTdh.
Using (Equation 12.10) to substitute for d P P in (Equation 12.4) leads to
(11.11)
dTdh=-k-1kgMR.
In the troposphere where the air is composed mostly of diatomic molecules k = 1.4, M = 28.88 which
yield
d T d h ≈ - 9.8 K / k m .
This value is greater (in absolute value) than the observed value quoted in the previous section of
- 6.5K/km. The difference is due to the fact that this ((simplistic” model neglects air moisture.
Observe that by integrating (Equation 12.11), one can compute P as a function of height using
(Equation 12.4).
11.2.2 Compressibility
In the previous chapter, we considered several aspects of Navier-Stokes equations (NSE) under the
assumption of incompressibility. However, in many applications (e.g., gas dynamics or atmospheric
applications), compressibility has to be taken into account. Boussinesq Approximation, in this
context, enables us to take into account some of these compressibility effects while retaining the
“flavor” of the incompressible equations (thereby reducing the nonlinearities in the equations). To this
end, we neglect the density variations in the continuity equation and the inertia term in the momentum
equations. However, these density variations do give rise to buoyancy forces in the momentum
equations.
The basic equations of the flow u = (u,v,w) are
∂ρ∂t+∇ ·(ρu)=0
∂u∂t+(u·∇ )u=-1ρ∇ p-gk
where p is the pressure, ρ is the density, k is a unit vector in the z direction, and g is the acceleration
due to gravity.
As a first step in this approximation, the mass continuity equation
∂ρ∂t+∇ ·(ρu)=0
is split into two equations
∇ ·u=0,∂ρ∂t+(u·∇ )ρ=0.
That is we consider the flow to be incompressible and ρ as a scalar that is carried over by the flow.
Furthermore, we assume that the density and pressure of the atmospheric fluid can be written as
ρ(x,t)=ρ0(z)+ρ′(x,t),p(x,t)=p0(z)+p′(x,t)
and
dp0(z)dz=-ρ0g
where ρ ′ ( x , t ) and p ′ ( x , t ) are small perturbations from ρ 0(z) and p(z) respectively. Introducing
the following definitions
P=pρ¯,N=-gρ-ρ0ρ¯,N02=-gρ¯dρ0(z)dz
where ρ ¯ is a constant ((reference density”, this leads to the following approximate system of
equations for the flow:
∇ ·u=0
∂u∂t+(u·∇ )u=-∇ P+Nk
∂N∂t+u·∇ N+N02w=0
where N 0 2 is called the Brunt‐Vaisala frequency. This system of equations yields a reasonable
approximation to the exact equations when the fluid is “almost incompressible.”
11.3 GENERAL CIRCULATION
The gross features of the atmospheric circulation are driven by convective currents due to differential
heating, Earth rotation, and the asymmetric distribution of land and sea. Due to these factors, the
atmosphere is divided into meridional and longitudinal convection cells.
To see how the meridional cells are formed, we observe that the equator receives much more heat
than the poles. Accordingly, we expect hot air to rise at the equator and travel in the upper
troposphere towards the poles, sink there, and then return along the surface to the equator. This global
picture was formulated by G. Hadley in the early 18th century and is named after him as “Hadley
circulation” (as a matter of fact such one-cell atmosphere exists on Venus).
On earth, the hot air rising at the equator sinks at about 30° (north and south) and thus forms the
tropical Hadley cell. At the descending branch of the Hadley cell, the air is dry. As a result, it creates
the great subtropical deserts on earth such as the Sahara and Gobi deserts.
A second meridional cell forms in the subtropics (between 30° and 60°). It is called the Ferrel
cell. Finally, there exists the polar cell in the arctic region.
The junction between the polar and Ferrel cells is called the polar front. Along this front in the
upper troposphere there is a strong band of westerly winds (i.e., winds blowing from the west) called
the jet stream. When the jet stream “meanders” south over the northern US, that region experiences
very cold weather.
We should note that due to the earth rotation, the winds in the upper troposphere of the Hadley cell
have a westerly direction. However, the returning surface wind is easterly (i.e., blowing from the
east). To understand this, observe that the tangential velocity of the earth is maximum at the equator.
For the Ferrel cell, this situation is reversed, and therefore we have “surface westerlies” in the mid-
latitudes. These are referred to as the “trade winds.”
Nonuniform heating due to the uneven distribution of land and sea drives the “zone overturning” or
Walker Circulation. In these cells, air rises at longitudes of heating (e.g., Indonesia) and sinks at
longitudes of cooling (west of South America). This circulation in normal years reinforces the
easterly trade wind across the equator. However, when this circulation reverses itself, it causes the
“El-Nino” current in the Pacific Ocean.
11.4 Climate
There exist various models for climate predictions in general and the computation of the mean
temperature of the earth. These models are usually classified by their degree of sophistication as 0-
dim, 1-dim, etc. The most sophisticated current models are the “Community Atmospheric Models”
(CAM) which were written at the National Center for Atmospheric Research (NCAR) [1].
In the 0-dimensional models, only the time dependence of the mean temperature is modeled and an
average is taken over the spatial dependence. For the one dimensional models, the dependence of the
temperature on the latitude and time is taken into consideration and so on.
In the following we provide a short narrative for Earth ‘‘climate modeling”
A key parameter in all these models is the albedo, which is the fraction of solar energy in the short
wave band which is reflected from Earth back into space. The value of the albedo depends on the
nature of the surface (ocean and different types of land, e.g., forests, deserts, etc.) and time (extent of
ice and snow cover). In the past, this “parameter” was modeled by various means, however, due to
recent advances in satellite imagery, it is now possible to compute the albedo accurately for each
location of the Earth.
In the following, we consider models with zero and one dimensions.
For the zero dimensional model, we take into consideration only the balance between the total
incoming and outgoing radiations which we denote by R and R o respectively.
(11.12)
CdTmdt=Ri-Ro
where T m is the global mean temperature, t is the time, and C is the heat capacity of the Earth system
(more precisely land, air, and oceans). To model the incoming and outgoing radiations, we let
(11.13)
Ri=Q{1-A(Tm)}
(11.14)
Ro=σg(Tm)Tm4.
Here, Q is the flux of the solar radiation, C is the heat capacity of the Earth system (more precisely
land, air, and oceans). A(T m ) is the mean albedo and g(T m ) is a “grayness-factor,” which measures
the deviation of Earth emissions from black body radiation due to the greenhouse effect.
Neglecting the spatial heat distribution on Earth, we can formulate a prototype model for the global
mean temperature (= climate). This can be done using the basic facts about the radiative balance of
Earth which were described in the previous sections. Thus, from Equations. (12.39)-(12.40) we infer
that
(11.15)
CdTmdt=Q[1-A(Tm)]-σ.g(Tm)Tm4
The factor g(T m ) was modeled by Sellers [2] by the following formula
(11.16)
g ( T m ) = 1 k tan h [ T m / T 6 ] , T 0 6 = 0.53 · 10 15 K 6
where k is the portion of Earth covered by clouds (under present conditions k ~ 0.5). In this model,
g(T m ) decreases as T m increases as the greenhouse effect becomes more pronounced.
As to the albedo, the following linear interpolation function was formulated by Sellers
(11.17)
A(T)=αM,T<T1αM-T-T1T2-T1(αM-αm),T1<T<T2αmT2<T
where a M and a m are the albedo values assigned to ice-covered and ice-free surfaces respectively (a
M = 0.85, a m = 0.25, T = 210°K, and T 2 = 275°K).
In a steady state, = 0 and (Equations 11.15) and (11.17) reduce to an algebraic equation which can
be solved for T m . We find then that this model has three equilibrium points, two of which are stable
while the third (in between) is unstable. The two stable equilibria correspond to “glaciation period”
and “present day” conditions.
Refinements to the model given by (Equations 11.15), (11.17) are obtained when we let Q depend
on a parameter A(t)
(11.18)
Q=λ(t)Q0
to take into account possible variations in the Sun radiative output with time.
We see that even this “prototype” model for Earth climate depends on many parameters whose
exact value is not known (and subject to change). This renders the predictions of this and more
sophisticated models somewhat unreliable with a large margin of error.
Many more elaborate models for the albedo have appeared in the literature. One of the major
sticking points in these models is the clouds that cover the Earth and their actual impact on the albedo.
For example, Bhattacharaya, et al. suggested a model for A(T m ) which is described by Fig. 11.1. In
this figure, the peak in the albedo near T = 220°K is attributed to the increased cloudiness near the
“ice-margin.” The use of this albedo model and (Equation 11.15) yield five steady state points, three
of which are stable.
A more detailed model for the albedo takes into account the different albedos of ocean, land, and
ice and the meridional extent of the ice cover of the earth. Thus,
(11.19)
A=aLβ+(1-β)aoc,aL=a1+a2M
where β is the land and ice percentage of the earth surface and a oc is the albedo of the ocean. The
albedo of the land (a L ) is composed of two parts: a 1 is the albedo of the ice free land for which M =
0, and the albedo of the ice sheet a 1 + a 2 M where M is the meridional extent of the ice sheet.
In a more refined model for the albedo of the land, M is a function of time and is governed by the
following differential equation
(11.20)
dMdt=λM-1/2[(1+ε(T))MT-M].
Here, MT is the meridional extent of the ice accumulation zone and e(T) is a ramp function. Thus, in
this formulation the earth climate is governed by two differential Equations, (11.20) and (11.20),
which depend on several parameters. As these parameters can vary, the climate can pass through
various bifurcations.
To introduce spatial dependence in these models, we have to change dT/dt into
(11.21)
DT/Dt=∂T∂t+(u·∇ )T
where u is the wind speed. However, as the climate time-scale is long (æ O(104yrs)), it is usual in
this context to eliminate u by applying the “eddy diffusivity approximation”
(11.22)
-(u·∇ )T≅ ∇ ·(νe∇ T)
where ve is the “eddy diffusivity coefficient.” With this approximation, (Equation 11.15) takes the
form
(11.23)
C(x)∂T∂t=QS(x){1-A(x,T)}-
σg(x,T)T4+∇ ·(νe∇ T)
where S(x) is the distribution of the solar flux on earth (if the earth axis had no tilt then S(x) = jcosø,
where φ is the latitude). To simplify, to some extent, one can assume that all quantities in (Equation
11.23) depend only on time and latitude. (Equation 11.23) then takes the following form:
(11.24)
C(ϕ)∂T∂t=QS(ϕ){1-A(ϕ,T)}-
σ g ( ϕ , T ) T 4 + 1 cos ϕ ∂ ∂ ϕ { ν e ( ϕ ) cos ϕ ∂ T ∂ ϕ }
We see that due to the extreme complexity of the climate system, there remains a lot of uncertainty
about our ability to predict the future climate of the earth. In particular, it is questionable whether
current climate models can reliably predict the impact of man-made inputs to this system.
Bibliography
[1] CCSM3.0 Community Atmosphere Model (CAM), http://www.ccsm.ucar.edu/models/atm-cam
[2] K. Bhattacharaya, et al.(1982) J. Atmos. Sci. 39 p. 1747-1773.
[3] J. Pedlosky - Geophysical Fluid Dynamics 2nd edition. Springer,NY.
[4] W.D. Sellers (1969) J. App. Met. 8 p. 392-400.
[5] J. Pedlosky - Geophysical Fluid Dynamics 2nd edition. Springer NY.
CHAPTER 12
Stochastic Modeling
CONTENTS
12.1 Introduction
12.2 Pure birth process
12.3 Kermackand mckendrick model
12.4 Queuing models
12.5 Markov chains
12.1 Introduction
In previous chapters, we considered the modeling of deterministic systems. For these systems,
information about the state of the system at time t determines with certainty its state at any later time.
For stochastic systems, on the other hand, no such certainty can be achieved, viz. the knowledge of the
state of the system at time t, enables us to predict only the probability that the system be in any of
several possible states in the future.
Our objective in this chapter is, therefore, to describe the basic logical steps that lead to such
stochastic models. To achieve this objective, we describe several stochastic models for various
growth and decay processes and compare to some extent their predictions with the corresponding
deterministic ones.
(12.3)
lim △ t → 0 p 1 ( △ t ) △ t = 0 i . e . p 1 ( △ t ) = O ( ( △ t ) 2 )
3. The probability p 2(t) that a tree reproduces more than once on [ t , t + △ t ] becomes
negligible, as ▵ t → 0, i.e.,
(12.4)
lim △ t → 0 p 2 ( △ t ) △ t = 0 o r p 2 ( t ) = O ( ( △ t ) 2 ) .
4. The probabilities of reproduction of a tree on two disjoint time intervals are independent of
each other.
Remark
A process that satisfies assumptions 1—4 or their equivalents is referred to as a “Poisson process.”
Mathematical Model:
Our basic objective here is to derive differential equations for the probability P N (t) that the tree
population at time t is equal to N.To begin with, we compute the probability that the tree population
will increase from N to N + 1 on the interval [ t , t + △ t ] . To do so, we observe that the population
will increase by one on [ t , t + △ t ] if one tree reproduces itself once and all the others do not
reproduce themselves on this time interval. Hence, since there are N possibilities to choose the tree
that reproduces itself, this probability is given by
(12.5)
P(N→ N+1)=
N(k△ t+p1(t))[1-(k△ t+p1(t))-p2(t)]n-1≈Nk△ t+p(△ t)
where p( ▵ t) = O(( ▵ t)2) . From this result, we infer that if the tree population at time t is N, then the
probability that the population remains unchanged at a later time is
(12.6)
PN(t+△ t)=PN(t)·(1-k△ tN)+O((△ t)2)
Similarly, P N (t + ▵ t), viz. the probability that the tree population at t + ▵ t is (exactly) N( ≥ N 0) is
the sum of:
1. The probability that at t the tree population is N and there was no reproduction on [ t , t + △
t].
2. The probability that at t the tree population is N - 1 and there was exactly one reproduction
on [ t , t + △ t ] (remember that the probability of more than one reproduction in ▵ t is
O(( ▵ t)2).
Hence,
(12.7)
PN(t+△ t)=PN(t)(1-kN△ t)+PN-1(t)k(N-1)△ t+O((△ t)2).
If the tree population at time t = 0 is N 0, (Equation 12.6) implies that it remains unchanged at time ▵ t
if
PN0(△ t)-PN0(0)=-PN0(0)k△ tN0.
Dividing this equation by ▵ t and letting ▵ t → 0, we obtain the following differential equation for the
tree population to be the same at time t
(12.8)
dPN0(t)dt=-kN0,PN0(0)=1
where we used the fact that P N 0 ( 0 ) = 1 .
Similarly, dividing (Equation 12.7) by ▵ t and taking the limit as ▵ t → 0, we obtain the following
differential equation
(12.9)
dPN(t)dt=k[(N-1)PN-1(t)-NPN(t)],PN(0)=0,N>N0.
(Equations 12.8) and (12.9) describe the stochastic process under consideration.
Analysis of the Model:
Since N is an integer, the system (Equations 12.8) and (12.9) can be solved recursively for N = N
0, N 0 + 1,... Thus, from (Equation 12.8) we conclude that
(12.10)
PN0(t)=e-kN0t.
Substituting this result in the differential equation for N 0 + 1, we obtain
(12.11)
dPN0+1dt+k(N0+1)PN0+1=kNoe-kN0t
whose solution is
(12.12)
PN0+1=N0e-kN0t(1-e-kt).
In general, however, it is possible (after a long algebra) to show that
(12.13)
PN(t)=(N-1)!(N-N0)!(N0-1)!e-kN0t(1-e-kt)N-N0.
Other related quantities which are important in the analysis of such models are the expected value of
the population size at time t
(12.14)
μ(t)=∑N=N0∞NPN(t)
and the variance
(12.15)
σ2(t)=∑N=N0∞(N-μ(t))2PN(t)
To evaluate μ(t), we differentiate (Equation 12.14) with respect to t and use (Equations 12.8), (12.9)
to obtain
(12.16)
dμdt=∑N=N0∞NdPNdt==-kN02PN0-k∑N>N0N2PN+k∑N>N0N(N-1
)PN-1
but
(12.17)
∑N>N0N(N-1)PN-1=∑N>N0(N-1)2PN-1+
∑N>N0(N-1)PN-1=∑N=N0∞N2PN+∑N=N0∞NPN.
Hence,
(12.18)
dμdt=kμ,μ(0)=N0
which leads to
(12.19)
μ=N0ekt.
Similarly, for σ 2 we obtain the differential equation
(12.20)
dσ2dt-2kσ2=kμ,σ2(0)=0
whose solution is
(12.21)
σ2(t)=N02e2kt(1-e-kt).
From Equation (12.18) we see that the deterministic version of this stochastic model deals only with
the expected value of the population size. This can be justified for large populations since then as t
becomes large
(12.22)
σ(t)μ(t)≈12N0≈0
viz. for large N 0 the probability distribution is sharply centered around μ(t).
Exercises
1. Build and solve a stochastic model for a pure death process (e.g., radioactive decay).
2. Build a model for a population in which both birth and death occur.
3. In our discussion of the pure birth process we assumed that k is constant. Discuss what
happens if k = k(N) or k = k(t) .
4. Derive and solve (Equation 12.20).
Evaluate μ(t) .
2. Evaluate P M-2(t) . Hint: use a computer algebra package.
3. A population of certain species consists of males and females. In a small colony, any male is
likely to mate with any female in any time interval of length ▵ t with probability
k ▵ t + O(( ▵ t)2) . Each such mating produces immediately one offspring which is equally
likely to be male or female. If M(t) and F(t) denote the number of males and females in the
population at time t, derive differential equations for P M,F (t).
A1. The probability that one customer arrives at the queue in [ t , t + △ t ] is k ▵ t + O(( ▵ t)2)
where k is a constant which is referred to as the mean arrival rate.
A2. The probability that more than one customer arrives to the queue in [ t , t + △ t ] is
O(( ▵ t)2) .
A3. If I 1 and I 2 are two disjoint time intervals, then the number of customers which arrive in I 1
does not affect the number of arrivals in I 2.
S1. If a customer is being serviced at time t, then the probability that the service is completed in
[ t , t + △ t ] is s ▵ t + O(( ▵ t)2) where s is a constant representing the mean service time.
S2. The probability that service to more than one customer is completed in [ t , t + △ t ] is
O(( ▵ t)2) .
S3. If I 1 and I 2 are two disjoint time intervals, then the number of customers whose service is
completed in I 1 does not affect the number of customers whose service is completed in I 2.
Remark
We note that by our assumption the probability of one arrival and one completion in [ t , t + △ t ] is
(12.30)
[k△ t+O((△ t)2)][s△ t+O((△ t)2)]=O((△ t)2)
Mathematical Model: To derive differential equations for P ℓ(t), we consider now the conditions
under which the queue length at t + ▵ t is ℓ > 0. These are:
1. The queue length at t is ℓ and there were no arrivals or departures during
[t,t+△t].
2. The queue length at t is ℓ - 1 and there was one arrival and no departures during [ t , t + △ t
].
3. The queue length at t is ℓ + 1 and there was one departure on [ t , t + △ t ] . Other possible
events are O(( ▵ t)2) by the assumptions and as noted in (Equation 12.30) Hence,
(12.31)
Pℓ(t+△ t)=Pℓ(t)(1-k△ t)(1-s△ t)+
Pℓ-1(t)(k△ t)(1-s△ t)+Pℓ+1(t)(s△ t)(1-k△ t)+O((△ t)2).
Dividing by ▵ t and letting ▵ t → 0 we obtain
(12.32)
dPℓdt=kPℓ-1+sPℓ+1-(k+s)Pℓ,ℓ>0.
Similarly, if ℓ = 0 then
(12.33)
P0(t+△ t)=P0(t)(1-k△ t)+P1(t)(s△ t)(1-k△ t)+O((△ t)2).
Hence,
(12.34)
dP0dt=-kP0+sP1.
Although this system of differential equations given by (Equations 12.32) and (12.34) can be solved
“in principle,” it is not possible to solve it recursively as we did in the previous sections. Hence, we
consider only the steady state solution for the queue, viz. the solution for P(t) when dP ℓ/dt = 0. Under
these conditions, (Equations 12.32) and (12.34) reduce to an algebraic system of equations:
(12.35)
-kP0+sP1=0
(12.36)
kPℓ-1+sPℓ+1-(k+s)Pℓ=0.
The solution of these equations is
(12.37)
Pℓ=ksℓP0.
The ratio q = k s is called the “traffic intensity” or the “utilization factor” of the queue.
Remarks:
If q > 1, then obviously the solution of (Equation 12.37) is meaningless since ∑ P ℓ = ∞
unless P 0 = 0.
If q < 1, then we must set ∑ P ℓ = 1 which yields P 0 = 1 - q and hence
Pℓ=qℓ(1-q).
For the expected queue length in the steady state, we obtain
(12.38)
E(ℓ)=∑ℓ=0∞ℓPℓ=∑ℓqℓ(1-q)=(1-q)q∑ℓqℓ-1=(1-q)qddq∑qℓ=(1-q)q(
1-q)2=q1-q.
Exercises
1. Build a modConsider a Markov chainel for N server queue viz. a bank with one line and N
tellers. Hint: Only S1 has to be modified as follows:
S1’: If customers are being served at time t, then the probability that service to at least one
of them is completed in [ t , t + △ t ] is
2. ℓs ▵ t + O(( ▵ t)2) if ℓ < N
3. 2s ▵ t + O(( ▵ t)2) if ℓ > N.
4. Compute the steady state solution and the expected queue length for the queue in exercise 1.
5. What happens if N = ∞ in exercise 1 (infinite server queue)?
Answers to Problems