Sei sulla pagina 1di 151

EBOOKS The Application of Mathematics

REEPING • REID
FOR THE in the Engineering Disciplines GENERAL ENGINEERING AND K-12
ENGINEERING David Reeping • Kenneth J. Reid ENGINEERING EDUCATION COLLECTION
LIBRARY This text serves as the companion text to Introductory John K. Estell and Kenneth J. Reid, Editors
Create your own Engineering Mathematics, which introduces common
mathematical concepts we see in engineering, including
Customized Content
trigonometry, calculus, and functions. This text assumes a
Bundle—the more level of mathematics of a high school senior, plus some ele-
books you buy, ments from the introductory text. Additional concepts we
the greater your
discount!
see in engineering are also introduced: specifically, matri-
ces, differential equations, and some introduction to series.
After, the basics of system analysis, including stability, are
The Application
of Mathematics

The Application of Mathematics in the Engineering Disciplines


explored.
THE CONTENT The concepts are presented using examples rather than
• Manufacturing strict mathematical derivation. As a result, this text likely will
not be an effective substitute for a differential equations

in the Engineering
Engineering
• Mechanical course, but by illustrating the implementation of differential
equations, it can be a companion to such a course. We pri-
& Chemical
marily use historical events as examples (including failures)

Disciplines
Engineering to illustrate the use of mathematics in engineering and the
• Materials Science intersections of the disciplines. We hope you develop an
& Engineering appreciation for how to apply these concepts, and find a
• Civil & new lens through which to view engineering successes (and
Environmental failures).
Engineering David Reeping is a PhD candidate in engineering edu-
• Advanced Energy cation at Virginia Tech and a National Science Founda-
Technologies tion graduate research fellow. David was the recipient of
the Remsburg Creativity Award for 2013 and the DeBow
Freed Award for outstanding leadership as an under-
THE TERMS
graduate student (sophomore) in 2014. He has extensive
• Perpetual access for experience in curriculum development in K-12 and cre-
a one time fee ates material for the Technology Student Association’s
• No subscriptions or annual TEAMS competition.
access fees Kenneth J. Reid is an associate professor in engineering
• Unlimited education at Virginia Tech. He earned his PhD in engi-
concurrent usage neering education in 2009. Among other awards, he and
• Downloadable PDFs his coauthors received the Wickenden award (2014), best
• Free MARC records paper award for the Educational Research and Meth-
ods Division of ASEE (2014) and IEEE-USA Professional
Achievement Award (2013) for developing the nation’s
For further information,
a free trial, or to order,
first BS degree in engineering education. He is active in David Reeping
engineering within K-12, including the Technology Stu-
contact: 
sales@momentumpress.net
dent Association (TSA) board of directors. Kenneth J. Reid
THE ­APPLICATION
OF ­MATHEMATICS IN
THE ­ENGINEERING
DISCIPLINES
THE ­APPLICATION
OF ­MATHEMATICS IN
THE ­ENGINEERING
DISCIPLINES
DAVID REEPING
KENNETH J. REID
The Application of Mathematics in the Engineering Disciplines

Copyright © Momentum Press®, LLC, 2018.

All rights reserved. No part of this publication may be reproduced, stored


in a retrieval system, or transmitted in any form or by any means—
electronic, mechanical, photocopy, recording, or any other—except for
brief quotations, not to exceed 250 words, without the prior permission
of the publisher.

First published in 2018 by


Momentum Press®, LLC
222 East 46th Street, New York, NY 10017
www.momentumpress.net

ISBN-13: 978-1-60650-907-4 (print)


ISBN-13: 978-1-60650-908-1 (e-book)

Momentum Press General Engineering and K-12 Engineering Education


Collection

Cover and interior design by S4Carlisle Publishing Service Ltd.


Chennai, India

First edition: 2018

10 9 8 7 6 5 4 3 2 1

Printed in the United States of America


Abstract

This text serves two purposes. First, it is the companion text to Introduc-
tory Engineering Mathematics within the General Engineering and K-12
Engineering Education book series. Introductory Engineering Mathemat-
ics introduces some of the math concepts we see often in engineering,
including useful trigonometry, calculus, and functions. This text assumes a
level of understanding of mathematics at the level of a high school senior,
plus a bit from the introductory text. The second purpose of this text is
to introduce additional useful mathematical concepts that we see in engi-
neering problem solving: specifically, matrices and differential equations.
The concepts are introduced by examples rather than strict mathematical
­derivation. Accordingly, the text probably wouldn’t be an effective substi-
tute for a course in differential equations, but it does intend to illustrate
how differential equations can be used and applied. Consider this text to be
a companion to such a course, or an introduction for someone interested in
exploring “why might differential equations be useful?” We use examples
drawn from engineering practice to illustrate the use of mathematics in
engineering, such as engineering failures both old and new. We hope you
see a broad coverage of mathematical concepts and develop an apprecia-
tion for how they apply to engineering. Perhaps this can give a new lens
through which to view engineering successes (and failures).

KEYWORDS

differential equations, engineering mathematics, failure analysis, m


­ atrices,
modeling, stability, systems, Tacoma Narrows Bridge
Contents

List of figures ix
List of tables xiii
Acknowledgments xv
Chapter 1 Modeling Systems in Engineering 1
Chapter 2 Differential Equations in Engineering 29
Chapter 3 Describing Systems Using Mathematics 55
Chapter 4 Analyzing Failure in Systems 81
About the Authors 127
Index 129
List of Figures

Figure 1.1. A network model of the major airports


in Washington DC. 4
Figure 1.2. A 2 × 4 board and a 2 × 4 matrix. 11
Figure 2.1. Initial conditions of a pendulum. 33
Figure 2.2. Process of solving a differential equation using
an integral transform. 40
Figure 2.3. An example of a pulse input. 44
Figure 3.1. Idealization of a rockslide. 56
Figure 3.2. Finding the output using the impulse response. 59
Figure 3.3. Example of a pole-zero plot using the complex plane;
in this instance, this function has three zeroes and five
poles (three of which are on the real axis). 61
Figure 3.4. Stability condition for systems. 62
Figure 3.5. Output of the stable system. 63
Figure 3.6. Pole-zero plot for the system. 63
Figure 3.7. Pole-zero plot for the hypothetical system. 64
Figure 3.8. Output of the unstable system. 64
Figure 3.9. Pole-zero plot for a marginally stable system. 65
Figure 3.10. Output of the marginally stable system. 65
Figure 3.11. A mass spring damper system. 66
Figure 3.12. Pole-zero plot for the mass spring damper
where k  b .68
Figure 3.13. Sample plot of the output where the dashpot
overpowers the spring. 69
Figure 3.14. Pole-zero plot for the mass spring damper where b  k .70
x  •   List of Figures

Figure 3.15. Sample plot of the output where the spring overpowers
the dashpot. 71
Figure 3.16. Pole-zero plot for the mass spring damper. 72
Figure 3.17. Sample plot of the output of the mass damper system. 73
Figure 3.18. Moving the poles closer and closer to the
unstable region. 74
Figure 3.19. Simple poles versus multiple poles on the
imaginary axis. 77
Figure 3.20. Cascade connection. 77
Figure 3.21. Parallel connection. 78
Figure 3.22. An example of a mix between cascade and parallel. 78
Figure 3.23. Implementing the systems in cascade connection. 79
Figure 4.1. Parabolic and hyperbolic representations of cables. 84
Figure 4.2. Diagram of the U10 plate and internal forces. 91
Figure 4.3. Free body diagram of the U10 gusset plate. 92
Figure 4.4. 3-4-5 Triangle. 92
Figure 4.5. Computing Cosine from 3-4-5 Triangle. 93
Figure 4.6. Finding the relevant forces to sum in the x direction
(in gray). 94
Figure 4.7. Finding the relevant forces to sum in the y direction
(in gray). 96
Figure 4.8. Demonstration of how a moment is calculated. 97
Figure 4.9. Illustration of Poisson’s ratio. 99
Figure 4.10. Stress–strain curve for a metal. 100
Figure 4.11. Illustration of shear 100
Figure 4.12. Forces acting upon a bolt in opposite directions. 102
Figure 4.13. Shear on the U10 plate. 102
Figure 4.14. An example of a distributed load. 105
Figure 4.15. Triangle created using the function f ( x ) = x on the
interval [0,1]. 106
Figure 4.16. Placement of the centroid. 108
Figure 4.17. A distributed load on a 6 ft long beam. 110
Figure 4.18. Resolution of a distributed load into a single force. 111
Figure 4.19. Breaking a load into pieces to simplify calculations. 111
Figure 4.20. Nonlinear distributed load on a 7 m beam 112
List of Figures  •   xi

Figure 4.21. Resolution of the nonlinear distributed load into


a single force. 113
Figure 4.22. Approximation of the integral using three trapezoids
(elements). 114
Figure 4.23. Comparison of sums of areas under the curve
(Steps 1 to 5 correspond to the steps in Table 4.4). 115
Figure 4.24. Examples of elements. 116
Figure 4.25. Shearing of a cylindrical object. 117
Figure 4.26. Shearing at the element level. 117
Figure 4.27. Idealized FEA solution. 118
Figure 4.28. AC voltage and the RMS, or effective, value. 119
Figure 4.29. Knee flexion angle. 123
Figure 4.30. Representation of the divot caused by degradation. 124
Figure 4.31. Identifying the maximum thickness. 125
Figure 4.32. An illustration of an “ill-structured” problem. 126
List of Tables

Table 1.1. Airport coordinates 4


Table 1.2. Evaluating the function’s derivatives at t = 1 6
t
Table 1.3. Evaluating e ’s derivatives at t = 0 7
Table 1.4. Summary of elementary row operations 13
Table 2.1. Forms of the solution for different situations 36
Table 2.2. Laplace transform pairs 42
Table 2.3. Common convolution pairs 52
Table 4.1. Shear strength 101
Table 4.2. Two simple special cases of distributed loads 109
Table 4.3. Comparison of convergence to the area under the curve
for different sums 115
Table 4.4. Comparison of patient’s flexion angle pre- and
postreplacement124
Table 4.5. Summary of materials common in knee replacements 125
Acknowledgments

We would like to extend our thanks to our wonderful colleagues and


friends who reviewed this book and provided invaluable feedback: Jeff
Connor, Alexandra Seda, and Lily Virgüez.
CHAPTER 1

Modeling Systems
in Engineering

If art can capture the human experience with a few brushstrokes,


­mathematical models capture the raw details of the same experience using
equations and formulas. Anything that we would quantify (make measur-
able) or explain using numbers is a candidate for modeling. Something
that is purely qualitative, meaning something that is not measured using
numbers, can also be mathematically modeled, strangely enough.
What is a model? A model is a representation of an item, a process,
an event, a person, etc. It may be a three-dimensional representation of a
physical structure, or it may be a simplified process meant to represent
a more complex one. Models are extraordinary helpful tools for cutting
costs and increasing the efficiency and effectiveness of our products and
processes. NASA certainly does not operate on a trial and error basis by
launching multimillion dollar space shuttles and hoping the journey is
successful. When the government is prepared to introduce a new program,
we may hear that the program is expected to save (or cost) so many ­billions
of dollars. To find this number, we can: either implement the program and
let it run for a few years and see what happens, or develop a model and
predict what we expect will happen. This is an example of a mathematical
model.
Why else would we want to make a model? Many fields of research
and businesses require a mathematical model to make decisions. These
can take the form of statistical models, an offshoot of mathematical mod-
els that involve statistics, where the main goal is to produce something that
can make accurate predictions. It is essentially the company genie that can
only grant one big wish and possibly do a few tricks.
More complex models are based off experimental data. After a cer-
tain number of trials, we may be pretty confident that you could come
2  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

up with a model to predict what will happen. That way, we can make an
accurate prediction of the results without having to collect more data.
This is particularly attractive once you start talking about outrageously
expensive, time-intensive massive experiments that may or may not be
feasible to run in real life. In the cases of real-world data, no perfect
model exists, we can only aim to create and improve the models to a
degree they are useful.

1.1  A NOTE ON CREATING MATHEMATICAL


MODELING

We’ll start nice and easy by creating and interpreting a common mathe-
matical model to make approximations of functions; however, the process
of generating a mathematical model generalizes to just about any model-
ing challenge. Much like learning a second language, there is a learning
curve in translating English to Math. Luckily, there are some tricks that we
can use to ease into these translations.

1. Read and understand the problem. This step should be obvious,


but we should always read a problem until we completely under-
stand what is asked. Not interpreting the problem correctly is where
most mistakes begin.
2. Once you understand the problem, pick out important infor-
mation. While this may be a more acquired skill, identifying given
information and necessary assumptions is crucial. Sometimes,
extraneous information—data or descriptions that are not relevant
or not helpful—may be included in the problem description. In the
real world, extraneous information is abundant and can take many
forms. Certain extraneous data could be completely meaningless
when the context of our goal is applied.
3. Think critically about the problem, make simplifying assump-
tions. Engineering and modeling are typically built around
open-ended problems—problems without a single correct answer—
which are not solvable unless assumptions are made. For instance,
the problems of “designing a vacuum cleaner” or “modeling the
traffic flow on the interstate” require some level of scoping to
become tractable. Further, engineering often utilizes a library of
messy formulas, most of which are interconnected. As we develop
a model or models, we should continue to examine if we are appro-
priately considering the given information, if our assumptions are
Modeling Systems in Engineering  •   3

valid and applicable, and that the work is progressing toward a use-
ful and meaningful solution(s).
4. Test your model. Once we go through all the trouble of creating
this work of art, we should determine if it works in the context
that you’re tasked to model. We should ask ourselves the following
questions:
a. Does our model match any known values given in the problem?
b. Do our predictions make sense?
c. What are we assuming to make this model work? Did we
­assume too much?

In the case of question (c), it is okay to make some assumptions—but


list and acknowledge them.

AN EXAMPLE OF A SIMPLE MODEL

For a simple example, let’s say we are in Washington DC and we want to


know the closest airport in terms of time. There are three major airports:
Dulles, Reagan, and Baltimore – Washington. Creating a mathematical model
of airport locations can take many forms, but the simplest would be a net-
work (also called a graph). The network will be made up of dots and lines
(nodes and edges) where each node is an airport (with one additional node as
our starting position) and each line connecting us to the airports is a “cost” to
move from one node to the next. In this case, we are concerned about the time
to travel to each airport, so the “cost” is the time to travel from the starting
position to the destination. Travel time can be easily calculated from

Distance = Rate × Time

If our starting position is given by the coordinates ( xS , yS ) and the air-


port’s coordinates are ( xA , yA ), then the distance is found using the usual
distance formula:

Distance = ( xS − xA )2 + ( yS − yA )
2

To find the distances, we could look up the latitude and longitude of


the airports and convert them to xy coordinates since latitude and longi-
tude are angular measurements, not linear like the flat xy plane. A quick
search using Google Maps yields the values given in Table 1.1.
4  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Table 1.1.  Airport coordinates

Airport Latitude (N) Longitude (W) X (km) Y (km)


Dulles 38.9531 77.4565 −36.4364 6.15

Baltimore 39.1774 76.6684 31.9339 31.049


Reagan 38.8512 77.0402 −0.321 −5.162

Note that X and Y will depend on what we choose as an origin. In this


case, we used the White House as the origin. Next, we can take our xy
coordinates and plot them in the plane to obtain Figure 1.1.
To calculate the costs along the edges (tDulles ,  tReagan ,  tBaltimore ), we will
take our distance formula and divide it by some average speed sAVG we
expect. This gives us a generic model for the travel time:

Distance from Start to Airport ( xA − xS ) + ( yA − yS )
2 2

t Airport = =
Average Speed sAVG

The complete mathematical model is the network as well, as it visu-


ally captures the relationship between our chosen starting point and the
resulting travel times to each of the airports.

Figure 1.1.  A network model of the major airports in Washington DC


(Note: the White House is at the origin).
Modeling Systems in Engineering  •   5

AN EXAMPLE OF A MORE COMPLEX MODEL

A model from calculus often used in engineering formulas to simplify our


work involves using an infinite series—typically Taylor and Maclaurin
series. For instance, we can use an infinite series to model a function—an
approximate of sorts. Power series serve as an overarching term to describe
any series modeling a function within a given interval, but typically, repre-
sentations need to be tied back to the geometric series somehow, especially
in a calculus course—which does not sound particularly useful. Instead,
we search for a more practical way to represent functions as series. Our
hunt for more useful techniques immediately leads us to Taylor series and
Maclaurin series. A Taylor series acts like an approximation of a function;
in fact, the same terminology is almost the same from the power series
discussion since a Taylor series is a power series.


fa (t ) = ∑cn (t − a)n
n=0

Note we said approximation; the old saying, “if it’s too good to be
true, it probably is,” certainly applies here. The disclaimer here is the
­“centered at a” portion, hence the subscript a in f a (t ).

WHAT DO WE MEAN BY “CENTERED AT a?”

The Taylor series is going to give us an approximation of a function as


a series of polynomial terms; eventually, if we had an infinite number
of these terms, we could have two functions that were exactly the same
(not just a good approximation). When we center our approximation
at a point a, we say that the function and the series approximating the
function are exactly the same at that point: the value of the function,
its derivative, 2nd derivative, etc. In general, the closer we are to point
a, the closer our approximation is to the function. In most cases, the
approximation is only valid for some small window around a.

The coefficients, represented by cn , are calculated by taking deriva-


tives and evaluating them at whatever the series is centered at, the a value:

f ( n) ( a )
cn =
n!
6  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Certainly the Taylor series gives us an approximation of the function,


but it is only valid on a particular interval. The interval in question is the
radius of convergence. We can find this interval using the ratio test.

1 cn + 1
= lim
r n → ∞ cn

1
Notice r is the radius of convergence, not .
r

Example 1.1: Finding a Taylor series for a simple function:

Before representing wildly special functions using a Taylor series, let’s


take a function that we want to express as a series around t = 1.

f ( t ) = 0.5t 3 + 0.5t 2 + 1

Not a wildly important or useful function by any means, but we will


use it to simply prove a point; all we need to do is take derivatives and plug
in 1 at each step to determine the coefficients (Table 1.2).

Table 1.2.  Evaluating the function’s derivatives at t = 1

Derivative Evaluated at 1

f (t ) , starting function 2

f ′ ( t ) = 1.5t 2 + t 2.5

f ″ (t ) = 3t + 1 4

f ″′ ( x ) = 3 3

f ″″ ( t ) = 0 0

 

0
f ( n) ( t ) = 0
Modeling Systems in Engineering  •   7

Following the format for the Taylor series, our center is 1—which is
the a value—and the coefficients are 2, 2.5, 4, 3, and 0 thereafter. This
means our Taylor series is,

 1  1  1  1
f1 ( t ) = ∑cn (t − 1)n = 2   + 2.5   ( t − 1) + 4   ( t − 1) + 3   ( t − 1)
 0!   1!   2! 
2
 3! 
3

n=0

f1 ( t ) = 2 + 2.5 ( t − 1) + 2 ( t − 1) + 0.5 ( t − 1)
2 3

***
A Maclaurin series may sound like it would be wildly different, but
it is only a special case of a Taylor series. Simply put, if we form a Taylor
series centered at zero ( a = 0 ), then it is a Maclaurin series—they are
practically synonyms.

Example 1.2: Finding the Maclaurin series for e t :

Arguably the most robust Maclaurin series we can find is a representation


for e t . The only requirement for us is to find the coefficients and the series
will take care of the rest.

y ( n) (0) n
et = ∑ n!
t
n=0

The process here involves taking the derivative of e x repeatedly and


plugging in zero for each derivative. Let’s make a table, even though we
should know better, and see what the coefficients should be (Table 1.3).

Table 1.3. Evaluating e t ’s derivatives at t = 0

Derivative Evaluated at 0

f (t ), starting function, = e t 1

f ′ (t ) = et 1

f ″ (t ) = et 1

f ″′ ( t ) = e t 1

 

f ( n) ( t ) = e t 1
8  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

No matter how many times we take the derivative of e t , the function


is ­going to stay the same. This means whichever derivative we evaluate at
zero will be e 0 = 1. While this certainly will not always happen for func-
tions in general, we can say in this case:

y ( n) ( 0 ) = 1     no matter what whole number n  may be

This means our Maclaurin series becomes



1
et = ∑ n! t n
n=0

Done! Now, where does this series represent e t ? In other words, what
is the radius of convergence? As engineers, it is crucial to know where
our approximations are valid; otherwise, mistakes are bound to happen.
Thankfully, we can easily find the radius using the ratio test for series:

1 cn + 1
= lim
r n → ∞ cn

Now if we substitute in the coefficients of the Maclaurin series with


the necessary adjustments,

1
1 ( n + 1)!
= lim
r n→ ∞ 1
n!

We do not need the absolute value signs, both quantities will always
be positive. Simplifying a bit gives us,
1 n!
= lim
r n → ∞ ( n + 1)!

Factorials are a little tricky to work with, but we can typically deal
with them by writing out what each factorial means:
1 n ( n − 1)( n − 2 )( 3)( 2 ) (1)
= lim
r n → ∞ ( n + 1)( n )( n − 1)( n − 2 ) (3)(2)(1)

From our last line, we can see almost all the terms will cancel, leaving
us with a fraction with a denominator heading toward infinity. This means,
1
lim =0
n→ ∞ n+1
Modeling Systems in Engineering  •   9

We are not quite done, we need to remember the equality from the
beginning:

1
=0
r

Ah, so the radius r needs to be extremely (infinitely) large to make this


equality remotely true. Therefore, the radius of convergence is infinity.
This would then imply, amazingly, that this representation of e t is valid
everywhere!
***

Example 1.3: Using a Taylor series to solve problems

Sure, representing a function as an infinite sum is a neat party trick, but


where is it useful? One immediate example is evaluating integrals which
have no antiderivative we can express using known elementary functions.
For instance, consider:

1
∫−1 e − t
2
dt

Since this integral has bounds, we know we just need to find the area.
Go ahead and try to integrate the function, you will make absolutely no
progress unless you use some advanced tricks. We have a trick of our own,
the Taylor series. Using the Taylor series for e t , we easily construct the
Taylor series for e − t .
2


1
et = ∑ n! t n
n=0

Now replace t by −t 2 .


1
∑ n! ( −t 2 )
n
e−t =
2

n=0

Done! Now we have Taylor series for e − t without putting forth much
2

­effort. Let’s replace the e − t in the integral with its Taylor series.
2


1
∫−1 n∑= 0 n! ( −t 2 )
1 1 n
∫−1 e − t
2
  dt =   dt
10  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

We can write out a few terms in the series to get an idea of what the
area may be:


1  1 1 1 
∫−1 n∑= 0 n! ( −t 2 )
1 n 1
  dt = ∫−1  1 − t 2 + 2 t 4 − 6 t 6 + 24 t 8 + …   dt

Now we integrate,

1  1 1 1 
∫−1  1 − t 2 + 2 t 4 − 6 t 6 + 24 t 8 + …   dt
1
 1 1 5 1 7 1 9 …
=  x − t3 + t − t + t + 
 3 10 42 216 
−1 

Evaluating at the bounds gives us,

 1 3 1 5 1 7 1 9 …
 1 − 1 + 1 − 1 + 1 + 

3 10 42 216
 1 1 1 1 
−  −1 − ( −1)3 + ( −1)5 − ( −1)7 + ( −1)9 + …
 3 10 42 216 

≈ 1.4950

The true area is calculated as approximately 1.4937, so we are surprisingly


close by only using a few terms.
***

1.2  USING MATRICES TO SOLVE PROBLEMS

When solving “big problems” with copious lines of equations in several


variables, the overwhelming feeling of dread may wash over those who
have not heard of the ubiquitous mathematical abstraction called the
matrix.
To begin, the definition of a matrix is the following:

Definition 1.1: Matrix—A rectangular array of numbers, symbols,


or e­xpressions arranged in n rows and m columns. When reporting a
­matrix’s size, we say it is n × m (n by m), just like pieces of wood (2 × 4s;
Figure 1.2).
Modeling Systems in Engineering  •   11

3 2 4 –1
about 2” 2×4 2 rows
–7 4 2 8

about 4” 4 columns

Figure 1.2.  A 2 × 4 board and a 2 × 4 matrix. (Note: a 2 × 4 board


actually measures 1.5 in. × 3.5 in.)

WHY IS A 2 × 4 NOT 2 in. × 4 in.?

When lumber is cut, the 2 × 4 board actually measures 2 in. by 4 in.


Lumber is dried in a kiln after it is cut and shrinks. After drying, it is
cut to the standard size for a 2 × 4: 1½ in. by 3½ in. Other board sizes
are similarly smaller than the “nominal dimensions.”

At first, matrices may sound like an unusual aside. What relationship


do systems of equations have with matrices? Let’s first establish what a
­matrix is. Say we have some system with a few equations, this system can
be transformed into a matrix:

 x1 + x2 + x3 = 1
  1 1 1 1 
 2 x1 + 2 x2 + 2 x3 = 2  2 2 2 2 
We can write this as a matrix!
 3x + 3x + 3x = 3  
 1 2 3 
 3 3 3 3 
   

How is this possible? Let’s break down how a matrix is set up. First,
each row corresponds to one equation. That means our first equation is
represented in the top row, the second goes below the first, the third below
the second, and so on.

 first equation   first equation 
  
 second equation To matrix form  second equation 
 third equation  third equation 
      

Now what about the columns? Well, the number of columns is depen-
dent on the number of variables we’re working with: the total number of
columns equals the number of variables on the left hand side of the equa-
tion plus column(s) for the solution(s). If an equation has four variables,
you’ll have four columns plus one column. Let’s take a look.
12  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

 3 x1 + 2 x2 + 5 x3 = 3  3 2 5 3 
  1 4 2 2 
 x1 + 4 x2 + 2 x3 = 2 To matrix form  
 x + 5x − 9 x = 7  1 5 −9 7 
 1 2 3
    x1 x2 x3 #

Easy enough, we just peel away the numbers in front of each variable
and place them in their correct spots in the matrix (remember to treat any
subtraction as the addition of a negative number). It’s easy enough to trans-
form a system into matrix form when all the variables are lined up with
each other (this makes the whole process much smoother). If a variable is
missing from one equation, but appears in another equation, place a 0 in
its spot in the matrix. For example,

 5 x1 + 2 x2 = 3  5 x1 + 2 x2 = 3  5 2 3 
 =  x1 missing in the second equation
 0 1 5 
x2 = 5 0 x1 + x2 = 5
   

Calculators and the computational software MATLAB love matrices; in


fact, the popular programming language MATLAB is short for ­“MATrix
LABoratory.” Provided you can arrange your problem statement as a
matrix, many software packages are there to come to your assistance.
If we want to feed a tasty matrix to MATLAB, all we need to do is
format the input as follows; A = [1, 2, 3; 5, 4, 3]. This assigns a 2 × 3
matrix with the values we picked to the variable A. Commas are used
to distinguish between the values on the same row and semicolons are
used to denote the next row.

Now, what do the numbers mean? Practically speaking, the numbers


contain the essence of the problem we are trying to solve. The values could
represent anything from the magnitude of forces to voltages at different
nodes in a circuit.
Never lose sight of the fact our matrices are meant to represent a sys-
tem of equations, so they have the same properties. For instance, the order
in which we write the equations in the system does not matter, so the rows
in the matrix can be swapped as many times as we would like without
changing the solution:

 Equation 1  Equation 2
 
 Equation 2 →  Equation 1
 Equation 3  Equation 3
 
Table 1.4.  Summary of elementary row operations

Elementary row operations


Multiply a row by a nonzero number Switch two rows Add a multiple of a row to another

2  1 1   2 2   1 1   2 2  −2  1 1   1 1 
  = 
−4  1 1   −4 −4 
↑  ↓  = >    1 1  =  1− 2 1− 2 
 2 2   1 1  ↓+    

We can multiply any row by a number other We can swap two rows at any point. We can multiply one row by a nonzero
than zero; it will not affect your answer at number and then add it to another row.
all.
Modeling Systems in Engineering  •   13
14  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

is the same as

 Coefficients of Equation 1   Coefficients of Equation 2 
   
 Coefficients of Equation 2  →  Coefficients of Equation 1 
 Coefficients of Equation 3   Coefficients of Equation 3 
   

We can also multiply or divide a row by any number:

“Number times Equation 1” is the same as “Number times Row 1”


“ 3  ×  ( x + 2 y − z   =  5)” is the same as “" 3  ×   [ 1  2   − 1  5 ] ”

Finally, we can add rows—or multiples of rows—together.

    Equation 1     Row 1
+  Equation 2 is the same as +  Row 2
=  New Equation 1 = New Row 1

        ( x + y + z = 1)          [1 1 1 1]
+  ( x + 2 y − z   =  5) is the same as +  [1 2 − 1 5]
=  2 x + 3 y + 0 z = 6 = [2 3 0 6]

Now, how do we find the solutions to equations using matrices? One


method is through Gauss–Jordan Elimination. Using the elementary
row operations (Table 1.4) we just talked about, we bring the matrix to
row echelon form to give us the solution(s) to the system.
We use the elementary row operations to bring the matrix to row
echelon form. We know the elementary row operations now, but what
does it mean for a matrix to be in row echelon form?

1. All nonzero rows (rows with at least one nonzero element) are
above any rows of all zeroes (all zero rows, if any, belong at the
bottom of the matrix).
2. The leading coefficient (which is the first nonzero number from
the left) of a nonzero row is always strictly to the right of the lead-
ing coefficient of the row above it.
3. All entries in a column below a leading entry are zeroes.

BONUS: If every leading coefficient is 1 and is the only nonzero entry


in its column, then the matrix is in reduced row echelon form.
Modeling Systems in Engineering  •   15

In a picture, the matrix will look something like the following:

 * # # 
 0 * # 
 
 0 0 * 

MATRIX MULTIPLICATION

Imagine we had a number in front of the matrix (we’ll call it n) like so:

 a b 
A = n 
 c d 

Just like distributing with parentheses, we can distribute the n to each


­entry in the matrix.

a b na nb
A =n =
c d nc nd

Multiplying two matrices is a bit trickier. When performing matrix


multiplication, we must remember that the number of columns in A
must be the same as the number of rows in B. If this condition isn’t met,
then the × multiplication isn’t valid.

m×n
n× p
AB = C m×p

(3 × 2) 1 5 3 rows
0 7
−1 3
2 columns

When we construct the matrix for the result of the multiplication,


we only care about the number of rows in A and the number of columns
16  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

in B. This combination will be the dimensions of the new matrix, the


original number of rows in A and columns in B do not matter after the
multiplication.

Example 1.4: Checking if Matrix Multiplication is valid

(1) Can we multiply a 2 × 3 and a 3 × 4 matrix?


(2) What is the indicator? The
number of columns in A must
( 2  ×  3)( 3  ×  4 ) = ( 2  ×  4 ) ? be the same as the number of
rows in B. Are these the same?
(3) ( 2  ×  3)( 3  ×  4 ) = ( 2 × 4) Since these two quantities are
the same, the multiplication
True. will work!

***
Now, how do we multiply? For this, we need to use dot product. Let’s
jump into an example and skip the theory so it is as basic as it can be.
Multiply a 2 × 2 matrix with a 2 × 2 and see what happens.

 1 3  0 1 
 2 5  5 6 
  

Look at this, we’ve taken the first row and considered the two col-
umns in the second matrix. What do we do with these?

 1 3  0 1 
 2 5  5 6 
  

Imagine taking the row and rotating it by taking the leftmost number
on top and stacking the other numbers below it. This stack should be the
same orientation as the columns.

1
1 3  → 
3

With these, we will multiply horizontally and then add the products
­together vertically. In other words, multiply what is in the bubbles, and add
the result of the bubbles:
Modeling Systems in Engineering  •   17

1 0
3 5

1  × 0 = 0
0 + 15 = 15
3 × 5 = 15

1 1
3 6

1  × 1 = 1
1 + 18 = 19
3 × 6 = 18

What do we do with the numbers we just calculated? These are entries


in the resulting matrix! How to know which is which depends on the col-
umn and row you considered. For example, we just took the first row and
first column, then the first row and second column. It only makes sense
that the result of those operations should go in the same places.

 15 19 
 ? ? 
 

Now we do that same thing for the second row, it will be the exact
same procedure.

 1 3  0 1 
 2 5  5 6 
  

Here, we will multiply across, add those results together, and place the
sum in the correct position in the matrix.

2 0
5 5

2  × 0 = 0
0 + 25 = 25
5 × 5 = 25
18  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

2 1
5 6

2  × 1 = 2
2 + 30 = 32
5 × 6 = 30

Finally, we have the product of the two matrices:

 1 3   0 1   15 19 
 2 5   5 6  =  25 32 
    

PROPERTIES OF MATRIX MULTIPLICATION

What happens if we switch the two matrices and try to multiply? Try this
problem

 0 1  1 3 
 5 6  2 5  = ?
  

How did we get a different result? Matrix multiplication is not com-


mutative. Even though regular multiplication works if we switch the order
that we multiply, the same cannot be said for matrices. Mathematically, if
we call the first matrix A and the second matrix B, then AB ≠ BA (the only
exception is if A = B ).
Although we can perform addition, subtraction, and multiplication on
matrices, matrix arithmetic does not support division. Instead, we multiply
by the inverse. Given a matrix A, the inverse of the matrix, A−1, is a special
matrix such that A and A−1 multiplied together produce an identity matrix.
For instance, if A is a 3 × 3 matrix, then:

 1 0 0 
AA −1
= A A= I =  0 1 0 
−1
 
 0 0 1 

Definition 1.2: Identity Matrix—An n by n matrix is called an identity


matrix if it has 1s along the main diagonal and 0s elsewhere. The notation
for the matrix is commonly written as I n, where n is the number of 1’s in
the diagonal.
Modeling Systems in Engineering  •   19

The identity matrix is essentially the matrix equivalent to the number 1.


Take a look:

I1 = [1]  1 0   1 0 0   1 0 0 0 
I2 =  
 0 1  I3 =  0 1 0   0 1 0 0 
  I4 =  
 0 0 1   0 0 1 0 
 0 0 0 1 

The hardcore mathematics, so-to-speak, is usually handled through


a scientific calculator or software since computation by hand is time-
consuming and tedious.
The simplest problem we may need to solve is

AX = B

where A and B are matrices containing constants and X is a matrix of the


variables for which we want to know the values. The equation could be
Ohm’s law (RI = V ), or Hooke’s law (KX = F ), but any problem sharing
a similar form will follow the same procedure. Using matrix arithmetic,
the symbolic solution is easily found:

X = A−1 B

but figuring out what A−1 B turns out to be can be quite tedious. This is
where technology comes in. After entering the known values of A and B,
the line of MATLAB code to retrieve the solution is just:

>> X = inv(A) × B

1.3 DETERMINANTS

The determinant is a scaling factor of a transformation computed using the


entries of a matrix, and we can use it to solve simultaneous equations. The
determinant pops up in calculus, linear algebra, engineering mechanics,
and so on. We can think of the determinant as an operator, which can be
written in the form det ( A). Let’s calculate det ( A):

We’ll call this 2 × 2 matrix, A

a b
A=
c d

det(A) = a b = ad − bc
c d
20  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

In the example above, the matrix is two by two, which has a conve-
nient formula for the determinant. How did we get it?

a b a b
−   
c d c d

We took the first diagonal including a and d, multiplied them to-


gether, and subtracted the second diagonal including b and c, with b and
c multiplied together. That’s all! What about larger matrices, such as those
with an extra column and row—three by three?

a b c
d e f
g h i

With a matrix bigger than two by two, we understandably have a much


larger formula. We could directly plug in numbers, but there’s little sense
in memorizing it:

a b c
d e f = aci − afh − bdi + bfg + cdh − ceg
g h i

Instead, we have a technique called row expansion. It is a quick


method of reducing the computation of the determinant of larger matrices
to computing the determinants of multiple 2 × 2 matrices, which is a
much simpler calculation. Before we walk through the method, we should
note that we can do the following process in any row or column since the
steps do not change. Now, here’s how we do it:

To start, we circle the first number in the top


row (this is the typical choice). Draw a line
a b c down the column and a line across the row
d e f (we don't need those numbers right now).
Box in the numbers that are left.
g h i
We do the same thing for the number in
a b c the next column.
d e f
g h i
Again for the last column.
a b c
d e f
g h i
Modeling Systems in Engineering  •   21

The boxed numbers will be their own matrix, so now we have these
determinants:

e f d f d e
h i g i g h
       

Then, the circled entries are multiplied by the corresponding determinant.

e f d f d e
a b c
h i g i g h
       

From here, there is a sign chart we need to follow depending on which


row or column we choose. Since we did row expansion across the top row,
we follow that part of the chart. That means, b will have a negative sign.

+ − +
− + −
+ − +

e f d f d e
+a −b +c
h i g i g h
       

Let’s put it all together in an expression for the determinant of the 3 × 3
matrix.

a b c
e f d f d e
d e f =a −b +c
h i g i g h
g h i

From here, we just use the equation for the determinant of a 2 × 2


matrix and apply it to each term, then simplify by distributing.

Example 1.5: Finding the Determinant of a 3 × 3 matrix

(1) Let’s find the


1 4 5
determinant of this
6 7 8
matrix using the row
2 3 2
expansion method.
22  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

(2) 1 4 5 We follow the procedure


6 7 8 we set up earlier. Circle
the first number, cross
2 3 2 out the numbers in the
1 4 5 same row and column,
6 7 8 then box in what’s left.
Move along the top
2 3 2
row and do the same
1 4 5 process.
6 7 8
2 3 2
(3) Take the boxed numbers
7 8 6 8 6 7
from each piece and we
3 2 2 2 2 3 have the following.
   
(4) 7 8 6 8 6 7 Now we add the circled
1 −4 5 numbers and put them
3 2 2 2 2 3
    with their respective
determinants, keeping
the sign chart in mind.
(5) 1 4 5 Now we can break up
7 8 6 8 the determinant into
6 7 8 =1 −4
3 2 2 2 three simplified pieces.
2 3 2
6 7
+5
2 3

(6) 7 8 6 8 6 7 Now we evaluate each


1 −4 +5 2 × 2 matrix.
3 2 2 2 2 3

= 1[( 7 )( 2 ) − ( 3)(8 )] − 4 [( 6 )( 2 ) − (8 )( 2 )]
+ 5[( 6 )( 3) − ( 7 )( 2 )]
(7) = 1[( 7 )( 2 ) − ( 3)(8 )] − 4 [( 6 )( 2 ) − (8 )( 2 )] Like before, it becomes
arithmetic.
+ 5[( 6 )( 3) − ( 7 )( 2 )]
= [14 − 24 ] − 4 [12 − 16 ] + 5[18 − 14 ]
=   −10 − 4 [ −4 ] + 5[ 4 ]
= −10 + 16 + 20
= 26 After all the
1 4 5 multiplication and
6 7 8 = 26 subtraction, we find
2 3 2 that the determinant
is 26.
***
Modeling Systems in Engineering  •   23

It is worth repeating that we are able to expand from any outside row
or column of the matrix as long as we follow the procedure (and the sign
chart!). While it is possible and “mathematically legal” to expand from
these other places, people usually tend to pick one spot and stick with it.

a b c a b c a b c
d e f d e f d e f
g h i g h i g h i

1.4  CRAMER’S RULE

We have one more tool to solve systems of equations. Unlike the matrix
method, we can solve for one variable at a time. Suppose we have the fol-
lowing system of equations:

 3x + 4 y + 7 z = 1

 2x + 5y + z = 3
 2 x + 4 y + 14 z = 0

Now, let’s set up the determinant of the matrix that goes along with it
and solve it, we’ll call it Δ.

3 4 7
∆= 2 5 1
2 4 14

5 1 2 1 2 5
=3 −4 +7
4 14 2 14 2 4

= 3( 70 − 4 ) − 4 ( 28 − 2 ) + 7 (8 − 10 )

∆ = 80

Once we know the value of the determinant, the solution to each piece
is represented by a formula, involving this special value, Δ.

∆x ∆y ∆z
x= y= z=
∆   ∆   ∆
24  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

From here, we need to define Δ  x; let’s set up the formula using these
­constants and involve them in the determinant:

 3x + 4 y + 7 z = 1

 2x + 5y + z = 3
 2 x + 4 y + 14 z = 0

In order to create this determinant, we need to replace the column


of x with the coefficients. This means our expression will look like this:

3 4 7
2 5 1
∆=
2 4 14
      x      y      z       

1 4 7
∆x = 3 5 1
0 4 14

Solving for Δ  x gives:

1 4 7
5 1 3 1 3 5
∆x = 3 5 1 =1 −4 +7
4 14 0 14 0 4
0 4 14

= ( 70 − 4 ) − 4 ( 42 ) + 7 (12 )

∆ x = −18

and we find:

∆x 18 9
x= =− = −
∆ 80 40

We can do the same procedure with y and z, the only change with the
­determinants is which row we replace with the coefficients.
Modeling Systems in Engineering  •   25

3 1 7 3 4 1
∆y = 2 3 1 ∆z = 3 5 3
2 0 14 2 4 0

The y column gets replaced. The z column gets replaced.

From here, we find:


∆y 58 29 ∆z 14 7
y= = = z= =− = −
∆ 80 40 ∆ 80 40

If we can sum up Cramer’s rule, what would we say?

1. Take all of the coefficients of each variable in the system and make
a determinant out of them. The columns will be the different vari-
ables and each row will be a different equation. This will be your
most important determinant, the Δ value.

 ax + by + cz = 0 a b c

 dx + ey + fz = 0         ∆ = d e f  
 gx + hy + iz = 0 g h i

2. Set up the Δ determinant for each variable by taking the free coef-
ficients (whatever each function is equal to) and replacing the
­variable’s column with those coefficients. This process is shown
with Δ x below:

 equation 1 = 1 1 b c

 equation 2 = 2           so,     ∆ x = 2 e f  
 equation 3 = 3 3 h i

3. Using the values for each Δ, use the easy formula to instantly get
the solution!

∆some variable
some variable =

26  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Example 1.6: Using Cramer’s rule

(1)  Let’s use Cramer’s


x+ y−z = 3
 rule to solve this
 −x − 3y + z = 2 system.
 2x − 2 y + z = 0

(2)  We need to
1x + 1 y − 1z = 3
 calculate the
 1x − 3 y + 1z = 2
− Δ for this
 2 x − 2 y − 1z = 0
 system. We
take all of these
1 1 −1 coefficients and
∆= −1 −3 1 set up the matrix.
2 −2 −1

(3) After row


1 1 −1
−3 1 −1 1 expansion and
∆= −1 −3 1 = −
−2 −1 2 −1 simplifying,
2 −2 −1
we find the Δ
−1 −3 for this matrix to

2 −2 be −2.

= [ 3 + 2 ] − [1 − 2 ] − [ 2 + 6 ] = 5 + 1 − 8 = −2

(4) We replace the x


3 1 −1
−3 1 2 1 column with the
∆x = 2 −3 1 = 3 −
−2 −1 0 −1 free coefficients
0 −2 −1
and use row
2 −3 expansion to

0 −2 find our ∆x,
which turns out
= 3[ 3 + 2 ] − [ −2 − 1] − [ −4 ] = 3[ 5] − [ −2 ] − [ −4 ] to be 21.
= 21

(5) Same procedure


1 3 −1
2 1 −1 1 with ∆y. Replac-
∆y = −1 2 1 = −3
0 −1 2 −1 ing the y column
2 0 −1
with the free
−1 2 coefficients and

2 0 expanding the
top row gives 5.
= [ −2 ] − 3[1 − 2 ] − [ −4 ] = [ −2 ] − 3[ −1] − [ −4 ]
=5
Modeling Systems in Engineering  •   27

(6) Replace the z


1 1 3
−3 2 −1 2 column with the
∆z = −1 −3 2 = −
−2 0 2 0 free coefficients
2 −2 0
and expand. The
−1 −3 result is 32.
+3
2 −2

= [ 4 ] − [ −4 ] + 3[ 2 + 6 ] = [ 4 ] − [ −4 ] + 3[8 ] = 32

(7) ∆x 21 We use our


x= = − equations to
∆ 2
quickly find the
∆y 5 solutions to the
y= = −
∆ 2 system.
∆z 32
z= =− = −16
∆ 2
***
Example 1.7: Using matrices to find voltages in a circuit

We would like to know the voltage drop across R7. Three equations are
needed to describe this circuit. Using nodal analysis, we can model the
circuit in terms of the three significant places that the voltage changes,
v1 ,  v2 , and v3. Since the drop occurs off of v3, that’s the voltage we seek.
We can use a technique called nodal analysis, a technique where we
write equations for the voltages at each node, then solve the system of
simultaneous equations. Note in the above example, the nodes are labeled
v1, v2, and v3. We won’t go into the theory behind finding these equations,
but using nodal analysis, we find the following:

 v1 − 12 v1 v1 − v2
 + + =0
 5 10 5
 v2 − v1 v2 − v3 v2
 + + =0
 5 5 10
 v3 − v2 v3
 + = 0 
 5 5
28  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

This system of equations (on the left, below) can be transformed into
a matrix by transferring the coefficients (bottom right):

 5v1 − 2 v2 = 24  5 −2 0 24 
  −2 5 −2 0 
 −2 v1 + 5v2 − 2 v3 = 0 To matrix form
  
− v2 + 2 v3 = 0  0 −1 2 0 

We know:

∆v1 ∆v2 ∆v3


v1 = v2 = v3 =
∆   ∆   ∆

We will simply use the three formulas to solve for each determinant:

∆ = 32;   ∆v1 = 192;   ∆v2 =  96;   ∆v3 =  48

Therefore, ∆v1 = 192 32 = 6V ;   

∆v2 = 96 32 = 3V ;  

∆v3 = 48 32 = 1.5V

***
The take-away message of matrices is that understanding matrices
conceptually is the key to knowing how to formulate the problem such
that software can understand it. While there is certainly a place for the
careful application of row operations, using them for practical problems is
usually not efficient.
CHAPTER 2

Differential Equations
in Engineering

INTRODUCTION

At the heart of an engineering system lies a characterizing equation, but


one that is unlike any ordinary equation we have encountered before. To
model systems with components varying with time, whether it is mechan-
ical or electrical, we need to implement the ideas of calculus in order to
form an equation that relates the input and the output. This relationship
is achieved through what is known as a differential equation. Unlike the
equations up to this point, which were algebraic, we are not looking for
numbers as solutions. Instead, we are looking for functions that satisfy the
relationship.
Definition 2.1: A differential equation is an equation that relates a func-
tion to its derivatives.

2.1  THE “TYPICAL” DIFFERENTIAL EQUATIONS

Although we are going to skip much of the theory of differential equa-


tions, we will be sufficiently prepared to discuss much of system modeling
by jumping to the main punch line of the course—solving equations of
the form

dn y d ( n   −  1) y … dy
Cn n
+ Cn   −   1 ( n   −  1)
+ + C1 + C0 y (t ) = x (t )
dt dt dt
30  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Since writing these equations using the full differential notation can be
tedious, we will use the more concise method of adding ′ to imply the deriv-
dy d2 y
ative of y; for example, we will denote as y ′ and 2
as y ″  or  y (2).
Rewriting the equation gives dt dt

Cn y ( n) (t ) + Cn   −  1 y ( n   −  1) (t ) + … + C1 y ′(t ) + C0 y (t ) = x (t )

This equation is a nonhomogeneous differential equation with con-


stant coefficients. Aside from being a mouthful to say, the equation carries
a considerable amount of information. The C0, C1, . . ., Cn−1 , and Cn terms
are just constants that could stand for a wide variety of physical measure-
ments like stiffness and inductance. The x (t ) and y (t ) terms are the input
and output of the system, respectively. On the left-hand side, the y ( n) (t ) is
the nth derivative of the output, y (t ); for example, y ( 3) ( t )   ( or   y ″′ ) means
the 3rd derivative of y (t ) or rather the cube of y (t ). Note that the right-
hand side could look just as complicated as the left-hand side, including
derivatives of the input, but is simplified for the sake of demonstration.
­Although the equation has quite a few moving parts, the solutions are
completely predictable. As the order of the equation rises (meaning the
highest derivative present, n), the complexity of the solution increases.
Yet, we can write the equation in a slightly different form to handle the
complexity:

(Some Derivative Operator)(Output) = (Some Derivative Operator)(Input)

Now, what do we mean by “some differential operator?”

Example 2.1: The typical differential equation form and the differential
operator

We think of the “derivative operator” loosely as a function where the vari-


able is the action of taking the derivative, which we will write as D. To
illustrate, consider the following differential equation. Again, x(t) is our
input, y(t) is our output—both of which are functions of time, t.

y ″′(t ) + 5 y ″ (t ) + y (t ) = x ″ (t ) − x (t )

Using our notation for the derivative, we can rewrite this as

D 3 { y (t )} + 5 D 2 { y (t )} + 1{ y (t )} = D 2 { x (t )} − 1{ x (t )}

We can factor out the common y on the left-hand side and x on the
right-hand side like any other variable or number to obtain
Differential Equations in Engineering   •   31

( D 3 + 5D 2 + 1) y(t ) = ( D 2 − 1) x(t )
( ) ( )
Both D 3 + 5 D 2 + 1 and D 2 − 1 can be called derivative operators.
***
The next step is to pretend we are solving a typical algebra problem
and find the roots of the derivative operator. From those roots, we will
have all the information we need for the entire solution. In this section,
we will examine two broad types of equations—homogeneous and non-
homogeneous. Arguably the simplest problem to solve is a differential
­equation that is homogeneous; we can easily spot them since the system is
not driven by an input, so the input x(t) will not appear.

(Some Derivative Operator)(Output) = 0

Definition 2.2: A differential equation is called homogeneous if the input


x(t) = 0.

To demonstrate, we will solve a simple homogeneous differential


equation using some intuition, the inspection method, which will lead us
to what “finding the roots of the differential operator” involves.

Example 2.2: Solutions to differential equations

One important distinction to make between algebraic and differential


equations is the way we describe solutions. When we solved something
like x 2 − 4 = 0, we could quickly say there were two solutions, 2 and −2,
using the quadratic formula or factoring. When we solve a differential
equation (assuming it is solvable), we cannot simply list all of the functions
that work. For the same reason integration needs an arbitrary constant, dif-
ferential equations can face the same issue of uniqueness in their solutions.
This means we need to write general solutions, which involve arbitrary
constants that paint the complete picture of the solution to the equation. If
we do not include any arbitrary constants, we have a particular solution.

Definition 2.3: The general solution is the solution of the differential


equation with arbitrary constants.
Definition 2.4: A particular solution is a specific solution of the differ-
ential equation with no arbitrary constants.
For instance, consider the below equation constructed for the sake of
illustrating the difference:

y″ − y = 0
32  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

This is a differential equation where y is the unknown function we


want to determine. Suppose we claim that y1 = e x and y2 = e − x are
solutions. How can we verify that? One method is brute force; plug both
of the supposed solutions into the equation and see if the equality holds.

( e x )″ − ( e x ) = 0

ex − ex = 0

Then, for y2 = e − x :

( e − x )″ − ( e − x ) = 0

e− x − e− x = 0

Both solutions work in the context of this equation, but y1 = e x and


y2 = e − x are both particular solutions. What if we added them together
and formed another solution, y3 = e x + e − x ? Even though this solution
incorporates both particular solutions, this is still a particular solution. To
make this a general solution, we need to add arbitrary constants that we
will call A and B.

yG = Ae x + Be − x

The constants add a bit of mystery, but also the sweet sense of gener-
ality. Plugging this solution into the differential equal reveals that A and B
could be anything:

( Ae x + Be − x )″ − ( Ae x + Be − x ) = 0

( Ae x ) ″ + ( Be − x ) ″ − ( Ae x + Be − x ) = 0
Ae x + Be − x − Ae x − Be − x = 0

Finding the general solution could be tricky or impossible through guess


and check alone, so we need an array of methods to solving these problems.

Example 2.3: Initial value problems

When finding the differential equation corresponding to a system, we


must take into account the initial state of the system. We will consider
Differential Equations in Engineering   •   33

the system to be the pendulum,


the input will be releasing the
arm from rest, and the output
will be the resulting movement
of the pendulum in terms of the
angle θ . Remember that the an-
gle is a function of time, so we
could write θ (t ), but this can be
redundant. In order to capture
what we call the initial state, we
need to consider how the system
can change.
In Figure 2.1, the pendulum
can start at any angle between Figure 2.1.  Initial conditions of a
0 and 180—up against the pendulum.
ceiling or slightly off-center.
­
Therefore, it would be wise to consider where the arm of the pendulum
begins before we release it—this will be the initial angle θ 0 . Second, we
could simply release the bob at the end of the arm or we could let it already
be traveling at some velocity v0 . From calculus and physics, we know that
θ (t ) is equivalent to the position while θ ′(t ) is the velocity (and θ ″(t )
would be acceleration). This means at t = 0, where we apply our input,
the initial conditions will be θ (0) and θ ′(0), respectively. In this case,
θ ( 0 ) = θ 0 and θ ′ ( 0 ) = v0. The differential equation of motion for a simple
pendulum is

g
 θ ″ + sin (θ ) = 0
l

where θ is the unknown function in terms of the angle. We can augment


this by incorporating the initial conditions we considered. Therefore, the
differential equation becomes

 g
 θ ″ + sin (θ ) = 0
 l
 θ (0) = θ0

 θ ′ ( 0 ) = v0

***
34  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Example 2.4: Solving a higher order equation using inspection, roots


of the differential operator

Let’s say we have some differential equation like

y ″ (t ) + 5 y ′(t ) + 6 y (t ) = 0

In this example, we are assuming no input is being supplied (which


means we find 0 on the right-hand side of the equation), so the equation
is homogeneous. As an educated guess, we will assume our favorite func-
tion, the exponential, is a solution. So,

Guess y (t ) = ce rt

where c and r are some numbers we do not know and t is the variable.
Remember, this is only an educated guess, so this will probably be a par-
ticular solution—not the entire solution. We do not necessarily need c, but
it would be fantastic to determine the value of r. To find r, we need to plug
our guess into the differential equation. First, we will find the first and
second derivative:

y ′(t ) = cre rt

y ″(t ) = cr 2 e rt

Now we can plug those in

y ″ (t ) + 5 y ′(t ) + 6 y (t ) = 0

cr 2 e rt + 5cre rt + 6 ce rt = 0

We can divide through this whole equation by ce rt since it will


never be zero (as long as c is not zero). Doing so reveals a fascinating
relationship:

r 2 + 5r + 6 = 0

This differential equation has a corresponding algebraic equation in


almost the exact same form! Remember the original differential equation?
Differential Equations in Engineering   •   35

y ″ (t ) + 5 y ′(t ) + 6 y (t ) = 0

D 2 { y (t )} + 5 D { y (t )} + 6 { y (t )} = 0

( D 2 + 5D + 6 ) y (t ) = 0
Compared to r 2 + 5r + 6 = 0, this is exactly the differential operator
where D is replaced with r! The values of r satisfying the equation are the
roots or the characteristic modes. Solving this equation tells us the r for
the exponential function, but there will be two solutions.

( r + 2 ) ( r + 3) = 0
r = −2 or r = −3

Since we have two solutions to the algebraic equation, both y1 = ce −2 t


and y2 = ce −3t sare solutions to the differential equation. In fact, there
will be as many algebraic solutions as the order of the differential equa-
tion. This class of differential equations allows writing the solution to be
­extraordinarily simple. In fact, if we add those two solutions together, we
will have another solution.
We can then introduce two different constants, we can call them A and
B, to point out that the two c’s could be different. This means, the general
solution is:

y (t ) = Ae −2 t + Be −3t

***
Depending on the roots of the differential equation, we could have
different “go to” solution structure. Our first example was the first case,
real and distinct roots—meaning all of our roots were different and none
of them contained imaginary numbers. The process to solving differential
equations of this type is routine and the subtlety lies in how the solution is
written; therefore, Table 2.1 can help guide us when determining the form
of the output.

Example 2.5: Mixing up the cases

Often the roots of the characteristic equation will be a combination of the


three cases. We demonstrate how easily we can use the table to write down
36  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Table 2.1.  Forms of the solution for different situations

The situation Form of the solution Example


(1) All roots are c1e r1t + c2 e r2t + … + cn e rnt r = 1, 2, 3
different and are c1e t + c2 e 2 t
real (real and + c3e 3t
distinct)
Roots are
r = r1 , r2 ,…, rn
(2) Roots are real, c1e r1t + c2 te r1t + … + cn t n e r1t r = 5, 5, 5
but some are c1e 5t + c2 te 5t
The first term is written
repeated + c3t 2 e 5t
as usual; if that term is
Roots are repeated, it gains a factor of
r = r1 , r1 , r1 ,…, r1 t for each repeat thereafter
(3) Roots are c1e at cos ( bt ) + c2 e at sin ( bt ) r = 3 ± 4i
complex c1e 3t cos ( 4t )
Roots are + c2 e 3t sin ( 4t )
r = a ± bi

the solution to these types of differential equations. Suppose we solved the


characteristic of the differential equation and there were a slew of roots:

r = 2,  2, 2,  6,  −4,  3 ± 2i,  3 ± 2i,  7 ± 2i .

In this scenario, we have all three cases at play, repeated roots that are
both real and complex! We can start with the simplest roots, 6 and −4,
these are the roots that are neither complex nor repeated.

yreal and not repeated ( t ) = c1e 6 t + c2 e −4 t

Next, we can consider the complex roots that are not repeated,
7 ± 2i ,

ycomplex and not repeated  ( t ) = c3e 7t cos ( 2t ) + c4 e 7t sin ( 2t )

The repeated real roots come next ( r = 2, 2, 2 )

yreal and repeated ( t ) = c5 e 2 t + c6 te 2 t + c7 t 2 e 2 t

Finally, the repeated complex roots can close out the solution
( 3 ± 2i,  3 ± 2i )
Differential Equations in Engineering   •   37

ycomplex and repeated ( t ) = c8 e 3t cos ( 2t ) + c9 e 3t sin ( 2t )


(
+ t c10 e 3t cos ( 2t ) + c11e 3t sin ( 2t ) )
Our complete solution is the sum of all four separate functions:

y =   yreal and not repeated ( t ) +   ycomplex and not rrepeated ( t ) +   yreal and repeated ( t )


+   ycomplex and repeated ( t ) 

=   c1e 6 t + c2 e −4 t +   c3e 7t cos ( 2t ) + c4 e 7t sin ( 2t )


+   c5 e 2 t + c6 te 2 t + c7 t 2 e 2 t +   c8 e 3t cos ( 2t ) + c9 e 3t sin ( 2t )
(
+ t c10 e 3t cos ( 2t ) + c11e 3t sin ( 2t ) )

HOW MANY ROOTS ARE POSSIBLE?

The number of roots possible brings us back to a formulation of the


Fundamental Theorem of Algebra. Since our roots are coming from
a characteristic equation, a polynomial of degree n, it must obey the
Fundamental Theorem by having n roots (counting duplicates).

***
The quality possessed by a homogeneous differential equation that
makes solving it so simple is the fact that we are assuming no input is
applied to the system. If the input x (t ) does not equal zero, then we have
the following nonhomogeneous differential equation:

(Some Derivative Operator )( Output ) = (Some Derivative Operator )( Input )

Definition 2.5: A differential equation is called nonhomogeneous if the


input x (t ) ≠ 0.
We will need to be cautious when solving nonhomogeneous equa-
tions, as there are few details to consider—we will illustrate this through
the next example.

Example 2.6: A Nonhomogeneous differential equation

Consider a system characterized by the following differential equation


where x ( t ) = t 2:

y ″(t ) − y (t ) = t 2
38  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

One subtle idea is to consider a particular solution y p (t ). This will


not be the entire solution, but it can give us some insight into how we can
crack this equation. Our challenge is to think about what kind of function
we could take the derivative of twice, subtract the original function, and
be left with t 2 ? Certainly it must be a polynomial; y p (t ) cannot be any-
thing like e t or sin(t ) or else we would see those on the right-hand side.
In order to form our particular solution, we keep the original term and take
the derivative of the right-hand side as many times as we can and add all
of them together.

{ }
t 2,    D t 2 = 2t ,   D {2t } = 2,   D {2} = 0

Therefore,

Sum of Right-Hand Side and Derivatives = t 2 + 2t + 2

We will make the following guess by inserting constants A, B, and C


as placeholders:

Guess y p (t ) = At 2 + Bt + C

Next, we need to find the second derivative of this function to find


the constants:

y ′p (t ) = 2At + B

y ″p (t ) = 2A

We can then plug y p and y ″p into the differential equation:

( 2 A ) − ( At 2 )
+ Bt + C = t 2

Simplifying slightly,

− At 2 − Bt + ( 2 A − C ) = t 2

Next we can compare coefficients, which gives us a simple system of


equations:
Differential Equations in Engineering   •   39

 for  t 2,    − A = 1

 for  t,       B = 0
 for constants,     2A − C = 0

From the system, we quickly notice A = −1, B = 0, and C = −2;


therefore,

y p (t ) = − t 2 − 2

We have our particular solution, but the problem is not quite solved
yet. The last subtlety to notice is the solution to the homogeneous version
of this equation can be built into the solution, this portion is called the com-
plementary solution, yc(t). To find the complementary solution, we pretend
we are solving the homogeneous version of the equation, which uses the
procedure we followed in the previous example. Here is the homogeneous
part of the equation:

y ″(t ) − y (t ) = 0

As before, we will factor out the differential operator, replace D with


r, and focus on the resulting algebraic equation:

D 2 { y (t )} − { y (t )} = 0

( D 2 − 1) y(t ) = 0
r2 − 1 = 0

By factoring, we know the proper values of r are ±1; therefore, the


complementary solution is

yc (t ) = c1e x + c2 e − x

To obtain the entire general solution, we add the complementary and


particular solution together:

y (t ) = yc (t ) + y p (t ) = c1e x + c2 e − x − t 2 − 2

***
40  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

2.2  INTEGRAL TRANSFORMS

We already examined the differential equations likely to appear in the


modeling of an engineering system. Although the solutions in the previ-
ous sections are perfectly valid, we can solve the same class of differential
equations using what is called an integral transform.
The name integral transform is quite literal. Using some integral,
which we have yet to specify, we can take a differential equation and
“transform it” into an equation that can be solved using algebra. We can
think of the integral transform as an operation like differentiation or typi-
cal integration, so consider a general integral transform, I { f ( t )} = F ( s ).
Notice two distinct things happened: (1) the transform calculated an an-
tiderivative (not too surprising) and (2) we swapped variables (from t to s).
The variable swap is what takes us to an equation that is solvable using
the standard ideas of algebra. Once we solve the algebraic equation, we
need to transform back to the world of differential equations. The inverse
integral transform is I −1 { F ( s )} = f (t ), which gives the solution to the
differential equation. Figure 2.2 outlines the general idea of how we get to
the solution using this integral transform.

Definition 2.6: An integral transform, I, is an operation acted upon a


function f (t ), in the form of an integral, which results in another function
as an output F ( s ). The transform has the following form:

I { f ( t )} = F ( s ) =
b b
∫a (A function of t  and s) f (t )  dt = ∫a K (t , s ) f (t )  dt
where a and b are real numbers. The bounds on the integral can be from
negative infinity to positive infinity (i.e., the entire real line).
We used the general notation I { f ( t )} because there are more breeds
of integral transforms than just one all-encompassing transform. For the

Too Integral
difficult Transform
Differential Algebraic
Equation Equation
Solve using
Solution Algebraic Algebra
Inverse Solution
Integral
Transform

Figure 2.2.  Process of solving a differential equation using an integral transform.


Differential Equations in Engineering   •   41

purposes of this text, we explore one transform that is important and


­common in engineering, the Laplace Transform.

2.3  THE LAPLACE TRANSFORM

The gentler, more vanilla, of the integral transforms is attributed to


Pierre-Simon Laplace. To set the stage, we consider the following integral
where t is the variable and s is a constant:


∫0 e − st   dt

We know that exponentials are easy to differentiate and integrate, so


this integral becomes


∞ 1
∫0 e − st   dt = − e − st  
s
0

To evaluate the antiderivative at the bounds, it could be useful to


1
­rewrite our result as − 1 by noticing e − st = st . Next, we consider the
se st e
bounds. Note we cannot “evaluate” anything infinite since it is not a num-
ber; we use a limit instead:


1  1   1   1
− =  lim − st  −  − s(0)  = 0 −  − 
se st  t →∞ se   se   s
0

The denominator of the first fraction is going to blow up to infinity, so


the entire fraction becomes zero. Plugging zero into the second equation
gives us − 1 . Thus, our integral is
s
∞ 1
∫0 e − st   dt = s

What is the lesson we can learn here? We started with an integral in


terms of a variable, t, and ended up with an expression in a new variable, s.
Is this an integral transform? Of course! We literally transformed an
42  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

integral equation into an algebraic equation. In fact, what we calculated is


the Laplace Transform of 1.

Pierre-Simon Laplace (1749–1827) was a French mathematician


who developed a series of equations to explain the tides of the ocean
and a method of using specific integrals to transform differential
equations, now known as the Laplace Transform.

Definition 2.7: The Laplace Transform of a function, given by L { f ( t )} ,


is defined to be the following:


L { f ( t )} = ∫0 f (t ) e − st   dt = F ( s )

where f (t ) is a function. F ( s ) is called the Laplace transform of f (t ).


We will not fret over the details of how certain functions transform
­(although they can be easily derived using the definition in the begin-
ning of the section), but it is useful to keep common transforms in mind.
Table 2.2 contains pairs of functions and their transforms that frequently
arise when solving differential equations.

Table 2.2.  Laplace transform pairs


Function, f(t) Transform, F(s)
δ (t ) 1

1
u(t )
s

n!
t n u(t ) (n is an integer)
sn

1
e −ω t u(t )
s+ω

ω
sin(ω t )u(t )
s2 + ω 2

s
cos(ω t )u(t )
s2 + ω 2

ω
sinh(ω t )u(t )
s2 − ω 2
Differential Equations in Engineering   •   43

Table 2.2.  Laplace transform pairs (continued )


Function, f(t) Transform, F(s)
s
cosh(ω t )u(t )
s2 − ω 2

ω
e − at sin(ω t )u(t )
( s + a )2 + ω 2

s+a
e − at cos(ω t )u(t )
( s + a )2 + ω 2

General Properties of the Laplace Transform

f ( t − a ) u(t − a ) e − as F ( s )

e − at f (t ) F ( s + a)

tf (t ) −F ′( s )

t aF ( as )
f  a
 

Note, two interesting functions are


shown in Table 2.2.
δ (t )—the impulse function or
delta function (or Dirac delta
function). The function has a value
of 0 except precisely at 0, where
the value is 1.

 0    t ≠ 0
δ (t ) =   
 1   t = 0

u(t ) —the unit step


function. The function is

 0    t < 0
u (t ) =   
 1   t ≥ 0

Our table is somewhat incomplete. We are working under the assump-


tion that our systems start up instantaneously and their outputs continue
indefinitely. Much like being caught in traffic or catching a flight, it is
44  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

( )= ( − 6) − ( − 10 )

Figure 2.3.  An example of a pulse input.

common to expect delays. We can also think about the idea of inputs that
only last a certain period of time, such as the input in Figure 2.3. These two
ideas are incorporated by including the unit step function and translations.
Like derivatives and integrals, the Laplace transform shares many same
properties—which contributes to why it is so useful. We stress the idea of
linearity. If we want to find L {af ( t ) + bg (t )} where f (t ) and g (t ) are
functions with constants a and b, then the Laplace transform of each func-
tion can be done separately and any constants can be factored out. Thus,

L {af ( t ) + bg ( t )} = aL { f ( t )} + bL { g ( t )} = aF ( s ) + bG ( s )

To find the original function f (t ) when we have the Laplace trans-


−1
form F ( s ) , we can apply the inverse Laplace transform L and use the
table to locate the correct f (t ).

L −1{ F ( s )} = f (t )

Example 2.7: A simple Laplace transform

Suppose we have worked toward solving for the voltage in a circuit, and
suppose we obtain the algebraic solution to the differential equation,

1
Y (s) =
1
s+
RC

where R and C are constants. In order to translate this algebraic solution


back to the time domain, we need to use the inverse Laplace transform;
therefore, we will apply L −1 to both sides.

 
 1 
L −1 {Y ( s )} = L −1 
1 
s + 
 RC 
Differential Equations in Engineering   •   45

Finding the inverse analytically can be tricky and is well outside the
scope of this text. It is more customary to use a transform table (Table 2.2)
provided earlier in the section. From the looks of the right-hand side, the
corresponding transform pair could be e −ω t u(t ):

1
e −ω t u(t )   
s+ω

where  means “transforms to.” To use a table to compute a transform,


we must first to do any legal algebraic manipulations to the expression
to make it look like one of the entries in the table. In this case, we have
almost no work to do,

1 1
=
1 s+ω
s+
RC

By examination, we can see ω = 1/ RC . Therefore, our transform is


simply

1
− t
y (t ) = e RC

***

Example 2.8: A more difficult transform

In conforming to the table, we may occasionally need to force a transform.


For instance, consider the following Laplace transform:

3
F (s) =
s5

We want to find the original signal f (t ) . The closest entry in the


table is

n!
t n u (t )   
sn

However, we can quickly verify that 3 is not 5 factorial. To force


the transform, we will factor out 3/5! from the numerator. Doing so
46  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

will leave us with a 5! as needed without changing the meaning of the


expression:

3 3  5! 
=  5
s5 5!  s 

We can think of this process as “unsimplifying” or “forced factoring”;


by undoing the cancelations and introducing the necessary values, we can
obtain the desired transform. Like before, we will equate the expression
with the proper skeleton:

3  5!  n!
 5  = n
5! s s

Therefore, n = 5. This means,

3 5
f (t ) = t u(t )
5!

***
Linearity is a wonderful property, but in order to fully transform a differ-
ential equation, we must determine what is meant by L { f ′ ( t )}; in fact,
{
what is the Laplace transform of the nth derivative  L f ( n ) ( t ) ? These }
questions are solved painlessly using integration by parts repeatedly. The
first three derivatives are as follows:

L { f ′ ( t )} = sF ( s ) − f (0)

L { f ″ ( t )} = s 2 F ( s ) − sf ( 0 ) − f ′(0)

L { f ″′ ( t )} = s 3 F ( s ) − s 2 f (0) − sf ′ ( 0 ) − f ″ (0)

The pattern is subtle, but it can be generalized to the nth derivative


of a function. Notice the pattern begins with the Laplace transform of the
function multiplied by s raised to the derivative of the function. Then, there
is a chain where the original function is evaluated at zero and multiplied
by a power of s. If we follow the mantra, “for the next term, reduce the
power of s by one, multiply it by the next derivative of the function at 0”
and stop when s is raised to the zero power, then we have the following:
Differential Equations in Engineering   •   47

{ }
L f ( n ) ( t ) = s n F ( s ) − s n −1 f ( 0 ) − s n − 2 f ′ ( 0 ) − s n − 3 f ″ ( 0 ) − … − sf ( n − 2 )
− f ( n −1) (0)

Using the definition of Laplace transform, linearity, and the transform


of a derivative, we can solve the following problem:

Example 2.9: A mechanical system

Consider the mass and spring system with an input force of x (which
is a function x (t )). If the mass is 1 kg and the spring constant is 1 N/m,
then the governing equation for this system is x = y ″ + y . Here, y cor-
responds to the position of the block, so the solution to this equation will
gives us a function for the position of the block at any time t. Assuming
this system is free of initial conditions, then y ( 0 ) = y ′ ( 0 ) = 0 . Now, let’s
take the Laplace transform of both sides:

L { x} = L { y ″ + y}

Between friends, we can say that y (or little y) corresponds to the


variable t and Y (big y) corresponds to the variable s. This would mean
L { x} = X and L { y} = Y .

X =  s 2Y − sy ( 0 ) − y ′ ( 0 )  + [Y ]

Since we have a system free of initial conditions,

X = s 2Y + Y

Now, our objective is to solve for Y so we can transform back to y,


the solution to the differential equation. We can factor out the Y on the
right-hand side and divide both sides by s 2 + 1 : ( )
(
X = s2 + 1 Y )
 1 
Y = X 2
 s + 1 

The solution to the differential equation is dependent on the input, X ,


as shown by our progress so far. Let us suppose the input is an impulse,
48  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

x = δ ( t ) ; this would be the equivalent to striking the block head on. We


know from the table that L {δ ( t )} = 1. This would mean, X = 1.

1
Y =
s +1
2

Take the Inverse Laplace Transform of both sides, and we see the
­position of the block will be given by some oscillation. In this case,

L −1 {Y } = y ( t ) = L −1 { }1
s2 + 1
= sin(t )

Since the mass is on a spring, this result should not be too surprising.
***

Example 2.10: An electrical system

Consider a circuit with a resistor, inductor, and capacitor (an RLC circuit)
with an input of x and the output y, which is the current. Say we have no
idea what the values of the resistor, inductor, or capacitor are; we will call
them R, L, and C, respectively. Using Kirchhoff’s voltage law, we have the
following differential equation:

1
Ly ″ + Ry ′ + y = x′
C

For simplicity, we’ll say the system has no initial conditions. Let’s
take the Laplace transform of both sides:

{
L Ly ″ + Ry ′ +
1
C }
y = L { x ′}

1
( )
L s 2Y − sy ( 0 ) − y ′ ( 0 ) + R ( sY − y ( 0 )) +
C
(Y ) = sX − x ( 0)

Using the assumption of zero initial conditions,

1
Ls 2Y + RsY + Y = sX
C
Differential Equations in Engineering   •   49

As we did before, let’s factor out the common Y term from the left-
hand side:

 2 1
 Ls + Rs +  Y = sX
C

Now, we will isolate the Y term:

 
 s 
Y = X
1
 Ls 2
+ Rs + 
C

This expression is going to cause some problems. For the sake of


simplicity, we will assume all of the constants are equal to 1; therefore,
L = R = C = 1. Note that this is not realistic in the physical world, but
we’ll set the values here to examine the form of the transform equation:

 s 
Y = 2 X
 s + s + 1 

Looking at the table, no common transform pair exists for this expres-
sion, but there are a few which are close like

s+a
e − at cos (ω t ) 
( s + a )2 + ω 2

We need to do some manipulations to make our situation match the


transform pair. Perhaps we can complete the square:

s s
= 2 2
s2 + s + 1  
1  1
s2 + s +   −   + 1
 2  2

This allows us to condense the fraction down to the form,

s s
2 2 = 2
 1  1  1 3
s2 + s +   −   + 1  s +  +
 2  2 2 4
50  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

but the transform pair calls for us to have an addition term in the n­ umerator
(+a). No problem, we can use the same technique of adding and sub-
tracting a term to obtain zero change in the expression, but will give us
the appearance of the transform we want. Since we need a “+1/2” in the
­numerator, let’s write “+1/2 − 1/2” in the numerator and split the fractions:

1 1
s s+
= 2 − 2
2 2 2
 1  3  1  3  1  3
 s +  +  s +  +  s +  +
2 4 2 4 2 4

Now we need to include another transformation pair,

ω
e − at sin (ω t ) u(t ) 
( s + a )2 + ω 2

We can equate our terms now; does the transform work?

1
s+ s+a
2 =
 1
2
3 ( s + a )2 + ω 2
 s +  +
2 4

Certainly! In this situation, a = 1 / 2 and ω 2 = 3 / 4 , so ω = 3 / 2.


Therefore, our first term has the following transform:

1
s+ −
t
 3 
2 e 2 cos  t u(t )
 2 
2
 1 3
 s +  +
2 4

Next, we will do the second term.

1
2 ω
=
 1
2
3 ( s + a )2 + ω 2
 s +  +
2 4

We have a little more work to do; we need to force a 3 / 2 in the


numerator (notice the numerator is ω , not ω 2 ) and factor out the 1/2.
Differential Equations in Engineering   •   51

To accommodate the 3 / 2 in the numerator, we need to multiply by


2 / 3 to avoid changing the expression’s meaning:

 3 
1 2    ω 
⋅  2  = constant  2
 ( s + a) + ω 
2 2
2 3 1 3
  s + 2  + 4 

We have the transform now!

 3 
1  2
 1 − 2t  3 
   e sin  t u(t )
 2 
2
3 1 3 3
  s + 2  + 4 

In our last step, we need to return to our original setup. By linearity,


we can say

1 1
s+  −t  3  1 − 2t  3  
2 − 2   e 2 cos  t − e sin  t    u(t )
 2  
2 2
 1 3  1 3 
  2  3
 s +  +  s +  +
2 4  2 4

Assuming our input was an impulse as before, x = δ ( t ); therefore,


X = 1, and we can conclude

 −t  3  1 − 2t  3 
y ( t ) =  e 2 cos  t − e sin  t  u(t )
  2  3  2  

***

2.4 CONVOLUTION

A minor but important detail to consider when using the Laplace trans-
form is the following common mistake:

L { f ( t ) g ( t )} = F ( s ) G ( s )  [ wrong !]
52  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

The Laplace transform of the product of two functions is not the


product of the Laplace transforms! It may be tempting, but we need to
define a new operation called convolution to fix the expression above.
Like the Laplace transform, convolution is defined as an integral and is
denoted by *.


f (t ) * g (t ) = ∫−∞ f (τ ) g (t − τ )  dτ

With this new operation, we can refine our incorrect equation by


­replacing multiplication with convolution like,

L { f ( t ) * g ( t )} = F ( s ) G ( s )  

To compute the convolution, it is desirable to use a table like the


­Laplace transform—like Table 2.3.

Table 2.3.  Common convolution pairs

First function Second function Convolution

f (t ) δ (t − T ) f (t − T )

u(t ) u(t ) tu(t )

u(t ) 1
e at u(t ) −
a
( )
1 − e at u(t )

1
e at u(t ) e bt u(t ) −
a−b
(
e at − e bt u(t ))
e at u(t ) e at u(t ) te at u(t )

t a u(t ) t b u(t ) a !b !
t a + b +1 u ( t )
( a + b + 1)!
Differential Equations in Engineering   •   53

While convolution is not multiplication, this operation does have


the same properties. For functions f ( t ), g (t ), and k (t ), the following
­statements hold:
Property Symbolic interpretation
Commutativity f (t ) * g (t ) = g (t ) * f (t )

f ( t ) * ( g ( t ) * k ( t )) = ( f ( t ) * g ( t )) * k ( t )
Associativity

f ( t ) * ( g ( t ) * k ( t )) = ( f ( t ) * g ( t )) + ( f ( t ) * k ( t ))
Distributive

Example 2.11: Computing the convolution using the properties


and table

Given the functions f ( t ) = t   u ( t ) and g ( t ) = t 2   u(t ), we can compute


f ( t ) * g (t ) using the table:

t   u ( t ) * t 2   u ( t ) 

In our table, this corresponds to the last entry:

a !b !
t a u (t ) * t b u (t ) = t a + b +1 u ( t )
( a + b + 1)!

In this case, a = 1 and b = 2

(1)!( 2 )!
t (1) u ( t ) * t ( 2 ) u ( t ) = t (1) + ( 2 ) +11u(t )
((1) + ( 2 ) + 1)!

After simplifying, we find

1 4
t   u (t ) * t 2   u (t ) = t u(t )
12

***
CHAPTER 3

Describing Systems Using


Mathematics

In the previous chapters, linear algebra and differential equations were


used to obtain the characterizing expression of the system. While s­ olving
differential equations, we used the Laplace transform to create a new
problem we could solve using algebra to find the function fitting the
­differential equation. With these tools, we can now close with arguably the
most important concepts in analyzing systems, the transfer function and
the impulse response.

3.1  THE TRANSFER FUNCTION AND THE IMPULSE


RESPONSE

To set the stage, we need to update our terminology regarding the


solutions of differential equations. In the traditional theory, solving a
nonhomogeneous equation required us to find a particular solution,
­
­determine the complementary solution, then add the two components
­together. While this method is functional, the Laplace transform allows us
to streamline our process in solving differential equations. Now, we can
say the solution to the differential equation can be broken down into two
pieces, a zero input solution and a zero state solution.

y ( t ) = zero input solution + zero state solution = yZI ( t ) + yZS ( t )

This result follows from the principle of superposition. For the


s­ ystems we are considering, those which are linear, we can look at each of
the effects influencing the result separately and simply add them together.
56  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Figure 3.1.  Idealization of a rockslide.

To visualize how this principle works, think of a rockslide; say a boulder


knocked itself loose and begins careening down a hill, picking up brush
and speed. Along the way, it strikes another boulder, which in turns strikes
another—dislodging other small rocks. In the aftermath, the boulders and
debris all roll into a nearby creek (Figure 3.1).
Nearby residents claimed it was the most nonviolent rockslide they
had ever seen, but a rockslide nonetheless. Even though the slide began
with a single boulder, the result was a culmination of what happened in
between. This shows that the large phenomenon “rockslide” is really the
sum of all of the individual rocks sliding.
The burning question now is how do we find yZI (t ) and yZS (t )? To
find the answer, we need to consider our general differential equation once
again:

(Some Derivative Operator )( Output ) = (Some Derivative Operator )( Input )

To avoid writing out “some derivative operator” repeatedly, we will


call them P( D ) and Q( D ), respectively. Now our general equation looks
like this:

P ( D ) y (t ) = Q ( D ) x (t )

Applying the Laplace transform to both sides will translate our differ-
ential equation to an algebraic equation. Due to the initial conditions on
y (t ), we will need to use the general formula for the Laplace transform of
a derivative. Note that this treatment is not needed for x ( t ) since it has no
initial conditions.

Q ( s )Y ( s ) − initial conditions ( s ) = P ( s ) X ( s )
Describing Systems Using Mathematics  •   57

While we will refrain from awkwardly writing the initial conditions,


do note that they will be a function of s as well. Once we have this alge-
braic equation, we can solve for Y ( s ) :

Q ( s )Y ( s ) = P ( s ) X ( s ) + initial conditions ( s )

Divide through by Q ( s ) and we have an expression for Y ( s ) :

P (s) initial conditions ( s )
Y (s) = X (s) +
Q (s) Q (s)

Our goal was to find the zero state and zero input response, and we
have succeeded with little effort. Simply observe the two terms that we
obtained:

P (s)
X ( s ) = component free from initial conditions,  input alone = yZS ( t )
Q (s)

initial conditions ( s )
= component driven by initial conditions,  no input = yZI ( t )
Q (s)

If our system is relaxed and has no initial conditions, then yZI ( t ) = 0


and we are left with

P (s)
Y (s) = X (s)
Q (s)

The fraction P ( s ) / Q ( s ) is monumentally important as it carries all


of the information about the original differential equation—which also
characterizes the system itself. We bestow a special title to this expression,
the transfer function:

P (s) Y (s)
H (s) = =
Q (s) X (s)

Provided with a transfer function, we can easily recover the ­differential


equation if needed.
58  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Example 3.1: All we need is a transfer function

Given the transfer function

s2 + 1
H (s) =
s 3 − 5s 2 + 1

we can find the corresponding differential equation by simply setting


H ( s ) equal to Y ( s ) / X ( s ) .

s2 + 1 Y (s)
H (s) = =
s − 5s 2 + 1
3
X (s)

Through cross multiplication, we can claim,

( s 3 − 5s 2 + 1)Y ( s ) = ( s 2 + 1) X ( s )
Take the Inverse Laplace Transform. This will turn each s into a
­derivative and transform Y ( s ) back to y (t ) and X ( s ) back to x (t ).

( D 3 − 5D 2 + 1) y (t ) = ( D 2 + 1) x (t )
Finally, distribute x (t ) and y (t ):

y ′′′ ( t ) − 5 y ′′ ( t ) + y ( t ) = x ′′ ( t ) + x ( t )

***
The transfer function H ( s ) is far more potent than the previous
e­ xample may have implied. We have been implicitly solving differential
equations by arriving at the following form:

Y (s) = H (s) X (s)

then finding the Inverse Laplace Transform yields,

y ( t ) = L−1 { H ( s ) X ( s )}

To simplify the right-hand side, we need to use a different operation


called convolution (again, denoted by *).

y (t ) = h (t ) * x (t )
Describing Systems Using Mathematics  •   59

System

( ) ( ) ( ) ( )

Impulse Response

Figure 3.2.  Finding the output using the impulse response.

The Inverse Laplace Transform of the transfer function has a ­special


name as well—we call it the impulse response. From the expression,
y ( t ) = h ( t ) * x ( t ), we can find the output for any input as long as we
know the impulse response. If the system is not relaxed, then h ( t ) * x ( t )
will yield the zero-state component (Figure 3.2).
Definition 3.1: A system is relaxed if the output is zero at time zero (t0 ).

Example 3.2: Finding the output of a system

Given the impulse response h ( t ) = e − t u ( t ) and the input x ( t ) = u ( t ), we


know the output of the relaxed system has the following form:

y (t ) = h(t ) * x (t ) = e − t u(t ) * u(t )

This can be located in the convolution table as

1 − e at
e at u ( t ) * u ( t ) = u (t )
−a

In this example, a = −1 .

1 − e ( −1)t
e ( −1)t u ( t ) * u ( t ) = u (t ) = 1 − e −t
− ( −1)

Therefore,

y (t ) = 1 − e −t

***
60  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

3.2 STABILITY

Balanced on the ends of long rods, the fine china plates spin with grace.
With the rods in hand, we keep each plate upright without falter. However,
an unexpected push from the side sends the plates into disarray—nearly
leaving their spots. We stumble and reorient the rods as we attempt to
­return to our original position, but was the push enough to have all the
plates eventually in pieces on the floor? If we consider ourselves to be
the system holding the plates up, can we be considered stable to avoid
the ­embarrassment of dropping the plates? The transfer function provides
insight into the stability of our system.
Definition 3.2: A system is called stable if the output of the system does
not diverge to positive or negative infinity.
Stability is intimately related to the so-called poles and zeroes of the trans-
fer function. Since H ( s ) is a function, we can find the zeroes and the
poles. While the zeroes are certainly important (solutions where our trans-
fer function is zero), the poles (solutions where our transfer function is
infinity) will determine the stability of our system.

Example 3.3: Poles and zeroes of the transfer function

Finding the set of poles and zeroes boils down to an algebra problem, so if
we have the following transfer function:

s+4
H (s) =
s2 + 9

For the zeroes, we set the numerator equal to zero and solve for s.

Numerator :  s + 4 = 0

This means s = −4 is a zero for this system. For the poles, we set the
denominator equal to zero and solve.

Denominator :  s 2 + 9 = 0

s = ±3i

In this situation, we have complex valued poles in the denom-


inator. While the appearance of imaginary numbers may have raised
Describing Systems Using Mathematics  •   61

e­ yebrows in the past, there is no need to be skeptical—this is a common


occurrence.
***
From the previous example, we stumbled upon complex numbers in the
denominator. To accommodate this, we need to realize that our input, s, is
a complex number. Therefore,

s = a + bi

for real numbers a and b. With this configuration, our poles and zeroes
now have graphical meaning and can tell us if the system is stable. When
dealing with complex numbers,
we create the complex plane
where the x-axis is redefined
to measure the real part of the
complex number and the y-axis
to measure the imaginary part. X O
To plot the poles and zeroes, we 3
will represent a pole by an X
O X
and a zero by an O. If multiple X O
poles or zeroes occur at a sin-
gle point, we will use a super-
script with the number of poles
or zeroes concentrated at that
Figure 3.3.  Example of a pole-zero
point. For example, if there are
plot using the complex plane; in this
n poles at a point in the complex instance, this function has three zeroes
plane, then we will write X n. and five poles (three of which are on the
An example of what a pole-zero real axis).
plot may look like is shown in
Figure 3.3.
Definition 3.3: A pole-zero plot is a tool for representing the transfer
function of a system in the complex plane.
From the pole-zero plot, we can answer a slew of questions about the sys-
tem, like “is this system stable” and “what is the transfer function of the
system?” Stability is a matter of how the poles are distributed in the com-
plex plane. In general, we consider a system to be stable if all of the poles
lie in the left-hand side of the complex plane. If there are poles on the
imaginary axis, then the system is marginally stable (if multiple poles are
62  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

stacked on top of one another,


then the system is u­nstable).
­Finally, if any poles exist in the
right-hand plane, then the sys-
tem is unstable (Figure 3.4).

Example 3.4: Stability in


systems compared with their
differential equations

Why is the left-hand side consid-


ered stable? Remember how we
STABLE UNSTABLE found the solutions to higher order
differential equations before the
Figure 3.4.  Stability condition for Laplace transform? If we had a
systems. system characterized by

y ″ (t ) + 5 y ′ (t ) + 6 y (t ) = x (t )

then we solved the corresponding algebraic equation by factoring out the


differential operator and replacing D with r.

( D 2 + 5D + 6 ) y (t ) = x (t )

r 2 + 5r + 6 = 0

( r + 2 ) ( r + 3) = 0

From this stage, we could use the resulting r values (r = −2 and


r = −3) as the exponents of our exponentials. If we assume our input is
an impulse, then the output will be the signal y1 ( t ) = c1e −2 t + c2 e −3t as
pictured in Figure 3.5.
The output gradually dies out and approaches zero after a certain
­period of time. What if we considered the system to be relaxed and solved
the differential equation using the Laplace transform instead? From the
point at which the differential operator was factored, take the Laplace
transform:

( s 2 + 5s + 6 ) Y ( s ) = X ( s )
Describing Systems Using Mathematics  •   63

Figure 3.5.  Output of the stable system.

Y (s) 1
= 2
X (s) s + 5s + 6

1
H (s) =
s 2 + 5s + 6

If we attempt to find our poles, then we end up with the exact same
solution:

s 2 + 5s + 6 = 0

where s = −2 and s = −3. In the


complex plane, these poles are
purely real, so they will lie in the
left-hand plane (Figure 3.6).
By our criteria, this system is X X
stable since both poles lie in the
−3 −2
left-hand plane. What happens
if we swapped the poles around
and said s = 2 and s = 3? Then,
clearly the system will be unstable
as shown in Figure 3.8. Figure 3.6.  Pole-zero plot for the system.
64  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Working backward, this means


the r values for the exponents are
r = 2 and r = 3 , respectively
(Figure 3.7); this corresponds to the
­output y2 ( t ) = c3e 2 t + c4 e 3t . The
X X plot of the output speaks ­volumes
2 3 as to why we would ­ consider it
­unstable (Figure 3.8).
Rather than decaying to zero,
this output races off to infinity—
quickly. What does it mean to
Figure 3.7.  Pole-zero plot for the be marginally stable then? For a
hypothetical system. ­system to be considered margin-
ally stable, we need poles on the
imaginary axis (Figure 3.9). To illustrate the point simplistically, will
choose s = ±i to be the poles.
According to our procedure for writing solutions to differential equa-
tions with complex eigenvalues, we use the structure of

c5 e at cos ( bt ) + c6 e at sin ( bt )

Our complex root is s = a + bi = 0 + (1) i , so a = 0 and b = 1.


Therefore, our marginally stable output is y3 ( t ) = c5 cos ( t ) + c6 sin ( t ).

Figure 3.8.  Output of the unstable system.


Describing Systems Using Mathematics  •   65

Figure 3.9.  Pole-zero plot for a


marginally stable system.

Figure 3.10.  Output of the marginally stable system.

In this scenario, the output (Figure 3.10) is neither dying out nor
­diverging, but any jolt to the system could cause it to become unstable!
***

Example 3.5: Analyzing system stability from start to finish

Say we have a mass spring damper system as shown in Figure 3.11. In


this system, we have a mass m, a spring k, and a dashpot (like a shock
absorber) b. The input will be the usual force setting the block into motion
and the spring will cause the mass to oscillate; however, the dashpot is
designed to resist motion by using viscous friction. Our question here is,
“is this system stable?”
Using Newton’s second law of motion,

∑ Forces  acting on an object = mass  ×  acceleration


66  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Figure 3.11.  A mass spring damper system.

Considering our output is the block’s motion in terms of position y (t ),


then the acceleration is y ″(t )—the second derivative of p­ osition. The force
inside the spring is dependent on the spring constant k multiplied by the
extension of the spring, which is related through the p­ osition of the block,
y (t ) . Finally, the dashpot’s force is dependent on the v­ elocity it is re-
sisting, y ′(t ), and the dampening coefficient b. Therefore, our ­differential
equation is

Force from dashpot + Force from spring + Input force


= mass of block  ×  acceleration of block

−by ′ ( t ) − ky ( t ) + x ( t ) = my ″ ( t )

while being mindful of directions (the spring and dashpot are resisting in
the negative direction). In the more standard form,

my ″ ( t ) + by ′ ( t ) + ky ( t ) = x ( t )

( mD 2 + bD + k ) y (t ) = x (t )

Assume the system is initially relaxed and take the Laplace transform,

( ms 2 + bs + k )Y ( s ) = X ( s )
Describing Systems Using Mathematics  •   67

Next, we will form the ratio of the output over the input—yielding the
transfer function. Moreover, we will assume the input is the unit impulse;
therefore, X ( s ) = 1.

Y (s) 1
H (s) = =
X ( s ) ms + bs + k
2

Now that we have the transfer function, we can find the poles and
zeroes. Looking at the numerator, this system does not have zeroes since
s terms do not even appear; instead, we head straight to the denominator
and set it equal to zero.

ms 2 + bs + k = 0

To avoid losing the purity in this example by yielding to phony values,


we will factorize this straight through. To complete the square, we need to
ensure the s 2 term has no constant (just the term s2)—divide through by m.

b k
s2 + s+ =0
m m

Add and subtract half of the square of the s term:

2 2
b  b   b  k
s2 + s+  −   + =0
m  2m 2m m

2 2
 b  k  b 
 s +  + −  =0
2m m  2m 

Now we can solve for s:

2 2
 b   b  k
 s +  =   −
2m  2m  m

2
b  b  k
s=− ±   −
2m  2m m
68  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

We can consider a few cases at this point. Suppose the resistance of the
dashpot dominates the stiffness of the spring, k  b . This would mean,

2
b  b  b b
s=− ±  =− ±
2m  2 m  2m 2m

b
Therefore, we have poles at 0 and − . Plotting these poles in the
m
complex plane yields a stable system (Figure 3.12).
We can find the output of the system with an overpowering dashpot.
Using the assumption k  b , the transfer function becomes H1 (we’ll
rename it as H1 to distinguish it from the original),

1 1
H1 ( s ) = =
ms + bs
2
s ( ms + b )

once we neglect the k and factor out an s in the denominator. To find the
output, we need to apply partial fractions to have any chance of using the
Laplace transform table—which is easily done in this case. We would find
the expansion to be as follows:

11 m 
H1 ( s ) =  − 
b  s ms + b 

By factoring out an m from the denominator of the second fraction,


we see the m’s will divide each other out.

X X

Figure 3.12.  Pole-zero plot for the


mass spring damper where k  b .
Describing Systems Using Mathematics  •   69

 
11 m  11 m 
 −  =  − 
b s ms + b b s  b 
m s +  
  m 

Doing so leaves us with

 
11 1 
H1 ( s ) =  −
b s s+ b
 
m

Taking the inverse Laplace transform of both sides by using the ta-
ble, then

1 − t
b
h1 ( t ) = 1 − e m  u (t )
b  

Plotting this signal reveals the displacement of the mass we would


expect. With the dashpot overwhelming the spring, the mass is not
­
­allowed to bounce back and forth. Therefore, the mass travels for a cer-
tain distance and stops completely (in Figure 3.13, the mass stopped after
traveling 0.01 m).

Block stops here

Figure 3.13.  Sample plot of the output where the dashpot overpowers the
spring.
70  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

What about the case where the spring dominates the dashpot, b  k ?
We have examined this system before; the problem reduces to the mass spring
system if b is approaching zero.

k k
s= ± − = ±i
m m

Plotting these poles in the


complex plane yields (Figure 3.14)
X To find the output in this case,
we return to the original transfer
function and apply our assumption.
For this case, we are assuming the
dashpot’s contribution is ­extremely
X −
small; therefore, our new transfer
function H 2 is

1
Figure 3.14.  Pole-zero plot for the H2 (s) =
ms + k
2
mass spring damper where b  k .

Before we can apply our ­table, we need to give this function a facelift.
We can force the transfer function to look like sin(t ), by factoring an m out
of the denominator:

1
H2 (s) =
 2 k
m s + 
 m

Recall the Laplace transform of sine,

ω
sin(ω t )u(t ) 
s2 + ω 2

k
We still need ω = on top! Easily done, we will forcibly factor the
m
numerator to bring out the term we want:

  k
1  1  m
= 
 k  k  m  s 2 + k 
m  s2 +   m  
 m m
Describing Systems Using Mathematics  •   71

With some radical arithmetic, we can use our table to take the follow-
ing fraction back to time domain:

 k 
1  m 
H2 (s) =  
km  s 2 + k 
 m 

The Laplace transform pair tells us the sine function will be our
­impulse response:

1  k 
h2 ( t ) = sin  t u(t )
km  m 

Like the previous situation, the output is precisely what we would


expect. If the spring overpowers the dashpot (Figure 3.15), then the mass
is going to oscillate as if the dashpot is not there.
We have looked at a few of the simplified cases, but what about the
transfer function we started with?

1
H (s) =
ms + bs + k
2

In the original case, we did most of the heavy lifting already. We know
the poles occur at

2
b  b  k b b 2 − 4 km
s=− ±   − =− ±  
2m  2m  m 2m m2

Figure 3.15.  Sample plot of the output where the spring overpowers the dashpot.
72  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

To achieve the full effect for this example, we will assume


b 2 − 4 km < 0 . Doing so allows us to coax an imaginary number out of the
radical. Note how the terms swapped to compensate for our assumption:

b 4 km − b 2
s=− ±i  
2m m2

Plotting the poles in the com-


plex plane yields a stable system
once again.
X Now that the poles are float-
ing off in the complex plane, how
does this affect our output? We can
− quickly figure out the output by
using the Laplace transform table
X − one last time. Back during the
process of finding the poles and
zeroes, we completed the square
Figure 3.16.  Pole-zero plot for the (Figure 3.16). Let’s put that form
mass spring damper. to use in the denominator.

1 1
H (s) = = 2 2
ms 2 + bs + k  b  k  b 
 s +  + −  
2m  m  2m 

The fraction looks suspiciously like the transform pair,

ω
e − at sin(ω t )u(t ) 
( s + a )2 + ω 2

However, we still need

2
k  b 
ω = − 
m  2m 
Describing Systems Using Mathematics  •   73

Figure 3.17.  Sample plot of the output of the mass damper system.

in the numerator of our fraction. Like before, we will forcibly factor the
numerator:

 k  b 
2 
 −  
1 1  m  2m  
2 2 = 2  2 2
 b  k  b  k  b   s + b  k  b 
 s +  + −  −    + −  
2m  m  2m  m  2 m    2m  m  2 m  

Applying the transform pair, we obtain the impulse response of the


system:

b  k 2
1 − t  b 
h (t ) = e 2m sin  t −   u(t )
k  b 
2  m  2 m  
− 
m  2m 

In the original system, we have a happy medium between our previous


two assumptions. When the dashpot overpowered the spring, the oscilla-
tion stopped almost instantly (Figure 3.17). On the other hand, our mass
bounced indefinitely when the spring overpowered the dashpot. This time,
the mass gradually comes to a halt at its starting position, which is often
the more desirable outcome (think of shock absorbers on a car).
***
From the previous example, it should be abundantly clear that pole place-
ment heavily influences how the system will behave. In the mass damper
74  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Poles moving right 1 2 3

X X X X

X X X X

1 2 3

Figure 3.18.  Moving the poles closer and closer to the unstable region.

system, we had poles floating in the complex plane. The more we move
the poles closer and closer to the imaginary axis and the right-hand side of
the plane, the more we flirt with instability. To illustrate, we will consider
similarly placed poles and gradually move them toward the ­imaginary axis
(Figure 3.18) for a pair of poles with the form s = a ± bi .
As the poles travel toward the right-hand side of the plane, the ­output
is snuffed out with less efficiency. At (1), the output is extinguished ­almost
instantly—how wonderfully stable! When the poles travel further, we
­obtain an output at (2) that takes a longer period of time to settle down.
Even closer at (3), the output begins to look like a pure oscillation. While
the signal will eventually be extinguished, it will take considerably more
time for every step closer we take to the right-hand plane. Once the poles
lie on the imaginary axis, there is little to stop the system from becoming
unstable; in fact, it will be marginally stable.
***

Example 3.6: Marginal stability and overlapping poles

Poles on the imaginary axis can be tricky to deal with. While the left-hand
plane is perfectly stable, we need to be mindful of multiple poles on the
imaginary axis. When solving differential equations the old-fashioned
way, certain situations would crop up where roots would be repeated. Put
Describing Systems Using Mathematics  •   75

simply, a repeated root would force us to multiply the root’s associated


function by a factor t for every repetition; for example, if −3 was the root,
then the associated function e −3t would become te −3t .
How does this affect stability? We can answer this question by consid-
ering the following system with no input:

( D 3 + 3D 2 + 3D + 1) y (t ) = 0
Going about our business by finding the roots to the differential
­operator yields,

s 3 + 3s 2 + 3s + 1 = 0

( s + 1)3 =0

This result points to three repeated roots of s = 1; therefore, the


­solution is

y ( t ) = c1e − t + c2 te − t + c3t 2 e − t

We can check if the signal settles down after a long period of time by
checking the function’s end behavior. This naïve test for stability manifests
itself as a limit:

t →∞
( ) t →∞
( ) t →∞
( ) t →∞
(
lim c1e − t + c2 te − t + c3t 2 e − t = lim c1e − t + lim c2 te − t + lim c3t 2 e − t )

In each case, we have an e − t term—which means we have a rapidly


increasing function in the denominator. The puny polynomials and con-
stants in the numerator are no match for the exponential growth in the
denominator; therefore, the output is eventually snuffed out.

( )
lim c1e − t + c2 te − t + c3t 2 e − t = 0
t →∞

This occurred in the left-hand plane, so of course we would have stability!


Marginal stability does not have the luxury of being able to handle ­overlapping
poles. To demonstrate, consider the following system with no input:

( D 4 + 2 D 2 + 1) y (t ) = 0
76  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Using the techniques we developed, we can conclude we have an over-


lapping set of poles at s = ±i . This means our output will look like the
following signal (in general):

y ( t ) = c1e at cos ( bt ) + c2 e at sin ( bt ) + c3te at cos ( bt ) + c4 te at sin ( bt )

where a is the real part of s and b is the imaginary part of s. For this
­example, a = 0 and b = 1 (check s = a + bi if this jump is unclear); thus,

y ( t ) = c1 cos ( t ) + c2 sin ( t ) + c3t cos ( t ) + c4 tsin ( t )

While e − t was our best friend last time, it did not receive an invite to
this party. On the imaginary axis, our friend is nowhere to be seen since
the real part of s is zero. Using the limit once again, we can quickly spot
the trouble. The extra factors of t are multiplying the cosine and sine terms,
so those terms will amplify as t becomes larger and larger. This divergent
behavior leads us to say,

lim ( c1 cos ( t ) + c2 sin ( t ) + c3t cos ( t ) + c4 t sin ( t )) → ∞


t →∞

which points toward instability.


***
To illustrate this in one picture, consider a pair of complex poles on the
imaginary axis in Figure 3.18. In the beginning, the poles are minding
their own business; however, we will move the outermost poles toward the
pair sitting closest to the origin. If we look at a moment before the poles
get acquainted with one another in (1), the output looks wonky—but it is
marginally stable. As soon as the poles meet at (2), we run into the issue
in our previous example. The extra factor of t causes the output to amplify
and oscillate without bound. Once the poles say their goodbyes at (3), the
output returns to being marginally stable (Figure 3.19).

3.3  CONNECTING MULTIPLE SYSTEMS


AND THE CONCEPT OF CONTROLS

We design systems to perform various duties, but real systems may not
quite perform the desired action or does so in an unsatisfactory manner.
Perhaps one system is not enough, and we need to break it down into
Describing Systems Using Mathematics  •   77

X Pole moving down

X 2
X
0
X
X
X
3

X Pole moving up

Figure 3.19.  Simple poles versus multiple poles on the imaginary axis.

multiple systems—how would we even connect them together? Even


worse, what if the system is the dreaded “u word”—unstable?

Example 3.7: System connections

Two simple types of connections exist for systems, cascade (also called
series) and parallel, which can be used to achieve different l­evels of
functionality. Consider three systems characterized by the transfer
­
­functions H ( s ), G ( s ), and K ( s ) . The three systems are said to be in a
­cascaded ­connection if we arrange them as in Figure 3.20.
In this configuration, the input is sent through the three systems one
after the other. Due to this, the output is dependent upon every subsys-
tem performing the required actions within the chain. The dependence is
­captured in the equivalent transfer function C ( s ), in which all of the trans-
fer functions are multiplied together:

C ( s ) = H ( s )G ( s ) K ( s )

A Rube Goldberg machine is one example of a system often


i­mplemented purely in cascade form. In this type of machine, a series
of ­contraptions (subsystems) form a chain reaction to perform a simple

( )= ( ) ( ) ( ) ( )=

Figure 3.20.  Cascade connection.


78  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

( )

( )= ( ) ( ) =

( )

Figure 3.21.  Parallel connection.

task in an overengineered and amusing fashion. If one of the stages in the


machine involves a makeshift catapult launching a walnut to knock over
a domino and the walnut misses the target, then successive tasks are not
performed—resulting in the desired output not occurring.
What if we used multiple makeshift catapults instead of one?
­Bombarding the domino with a hail of walnuts will surely increase the
chances of the chain reaction continuing. In terms of system connections,
the ­implementation of multiple systems simultaneously is called parallel
­connection (Figure 3.21).
Rather than going through each system one at a time, the input is
sent through each system at once. In this case, the systems are indepen-
dent of each other. If the output only requires a subset of the systems to
complete their duties, then it would be alright for one or more of the sys-
tems to fail. Like the cascade connection, the independence can be seen
in the ­equivalent transfer function P( s )—the transfer functions are added
together,

P (s) = H (s) + G (s) + K (s)

We can combine these types of connections to form a slew of different


configurations like the one shown in Figure 3.22.

( )
( )= ( ) ( )=

( )

Figure 3.22.  An example of a mix between cascade and parallel.


Describing Systems Using Mathematics  •   79

The transfer function for the system above R ( s ) will be

R ( s ) = K ( s ) [ H ( s ) + G ( s )]

***

Example 3.8: Stabilization?

Armed with the idea of parallel and cascade connections, can we stabilize
a system? Consider the following transfer function:

1
H (s) =
( s + 1)( s − 1)

Clearly H ( s ) is unstable since one of the poles is in the right-hand


plane, s = 1, but what if we could simply “remove” the pole at s = 1?
Mathematically, this would involve introducing a factor of ( s − 1) . ­Luckily,
we can use a cascade connection to do so! Let’s define a new system Q( s )
using the factor we need to cancel:

Q (s) = s − 1

Now, we can place the two systems in cascade connection (Figure 3.23).

( )= ( )=

Figure 3.23.  Implementing the systems in cascade connection.

The equivalent system will be called H stable ( s ) and is found by multi-


plying H ( s ) and Q( s ) together.

 1  1
H stable ( s ) = H ( s ) Q ( s ) =   ( s − 1) =
 ( s + 1) ( s − 1)  s +1

As desired, the unwanted pole is gone and the system is now


­stabilized! Well, at least mathematically. In practice, it can be difficult to
80  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

stabilize systems this way since we would need to physically implement


these transfer functions as an electrical or mechanical device. Once we
begin to play with hardware, manufacturing and rounding errors emerge
from the cracks to cause us headaches.
In the real world, our mathematical models give us perfect results:
we ­assume all components are at ideal values, and all processes work
flawlessly. However, electrical and mechanical components have flaws
or tolerances, which can make our systems become unstable or behave
in a slightly different way than our perfect models. When working with
­models, we should always do a “reality check”—if our components have a
tolerance of ±5 percent, will our model still hold true?
***
Rather, we turn to control engineering to inform our designs or make
­adjustments to the system in real time without trying to stabilize it as in
the example.
CHAPTER 4

Analyzing Failure
in Systems

4.1  “GALLOPING GERTIE,” THE TACOMA


NARROWS BRIDGE

Back in the 1940s, the Tacoma Narrows Bridge in Washington was hailed
as the third largest suspension bridge in the world. This bridge regularly
serves as the center of a specific subject in the field of science, mathe-
matics, and engineering. Surely because of the marvel of its incredible
length, right? Not quite. A tip-off to this bridge’s fascinating behavior can
be picked up from its nickname, “Galloping Gertie.” The bridge was given
this title when the construction workers noticed that the deck would yield
to wind conditions and bend up and down. Despite this, the bridge was still
opened to the public on July 1, 1940.
Fast-forward a mere 4 months to November 4, 1940—a fateful, windy
day had arrived. That vertical movement the bridge displayed since open-
ing day transitioned to violent twisting—called torsion. This behavior
lasted roughly 45 minutes until the bridge failed and collapsed into the
river below. The Tacoma Narrows Bridge was replaced with a much more
stable design within a few years.

FIXING THE TACOMA NARROWS BRIDGE

After the bridge opened and it was found to sway, cars would line
up and pay 55 cents (plus 15 cents per extra passenger) to “ride”
the bridge. Engineers designed a scale model and investigated
­anchor cables. When anchor cables were attached to the bridge, they

(continued )
82  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

snapped in the wind. Since the solid nature of the bridge meant that
wind could not pass through, proposals to drill holes through the
bridge or attach semicircular deflectors were received. The proposal
to attach deflectors was approved by the Washington Toll Bridge
Authority in November 1940. The bridge collapsed on November 7,
1940, before the deflectors could be installed.1

Mathematicians and engineers are mainly interested in modeling


the  vertical motion and torsion, all in hopes to determine exactly why
the  bridge failed. The most common base model used is a differential
equation21:

This large term


describes the cable
resistance.

− cos( ) sin( )

This is the second


derivative of the This term models the Finally, this term is
angle rotation. bridge’s wind the external force.
resistance.

From this model, we can use MATLAB to solve this equation and plot
the solutions. We’re not as concerned as solving this differential equation.
Instead, we could to model the bridge by using a free body diagram ap-
proach, including the forces, then perform some calculations.
Definition 4.1: A free body diagram is a drawing of a component
showing each force acting upon the component at the location where
each acts.
Any bystander could look at the Tacoma Narrows Bridge as it twisted
and inferred the violent behavior would surely cause the bridge to fail.
Still, it’s not that easy. Various theories exist as to why the bridge ulti-
mately fell apart. Looking at the picture of the bridge and referencing our
free body diagrams, we can start to make some inferences. Let’s explore
some of the theories. . .

1
History of the Tacoma Narrows Bridge. http://www.lib.washington.edu/specialcollections/
collections/exhibits/tnb/opening
2
P.J. McKenna. 1999. “Large Torsional Oscillations in Suspension Bridges Revisited: Fixing
an Old Approximation,” The American Mathematical Monthly 106, pp. 1–18.
Analyzing Failure in Systems   •   83

SIMPLE HARMONIC MOTION

When the deck of the bridge was in torsion and yielding due to the wind,
the movement resembled simple harmonic motion. The equation itself can
be represented as a simple second order differential equation:

d2 y
m = − ky
dt 2

where m is the inertial mass of whatever’s moving, y is its displacement


from its starting position, and k is the spring constant. Using some in-
tuition when solving the equation above, we find that the solution is a
sinusoidal function.

Angular Frequency—the rate of


Amplitude—the biggest distance the change in rotation (in radians per
object goes from its normal position. second).

Phase—how much the function


( ) is offset from the usual sine and
cosine deal.

The solution is fairly simplified to make it a little more user-friendly,


we collapse some of the complicated expressions and package them un-
der a symbol. For example, the angular frequency is represented by the
Greek letter omega—it contains another quantity, frequency, f.
The frequency is just the number of cycles per time, that’s measured
in hertz (Hz).

#  of times an event occurred
ω = 2π f ,   where   f =
timee

MODELING THE CABLES OF THE SUSPENSION BRIDGE

The cables are theorized to be one of the culprits that contributed to the
Tacoma Narrows Bridge’s failure. With the violent torsion caused by the
wind, the cables would yield with the bridge, causing a large internal
force—tension. To begin looking at the cables, we first need to model or
“mathematize them,” meaning devise a mathematical model to describe
84  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

their shape. The shape that the cables create looks like a parabola, but it
can also be a catenary—which is the shape of an idealized hanging chain
(almost like a parabola). We won’t derive the equations, but here are the
two representations of the cables—one is parabolic and the other is hyper-
bolic. Below that is what those would look like in the first quadrant:

1   x  
y1 = x2 y2 = 3, 545.708  cosh   − 1
7,000     3, 545.708  

Feet
4,000

1
=
3,000 7,000

2,000

= 3,545.708 cosh −1
3,545.708

1,000

0
Feet
0 1,000 2,000 3,000 4,000

Figure 4.1.  Parabolic and hyperbolic representations of cables.

The length of the largest span was 2,800 ft. We’d like to estimate the
length of the cable to get some more information about the longest section
of the cable system. Doing so requires the following integral:

1 + [ y ′ ( x )]   dx
b 2
Length = ∫a
For the parabolic model, we need to calculate y1 ′(t ). Since the func-
tion is a parabola, it is simply the power rule:

2
y1 ′ ( t ) = x
7,000
Analyzing Failure in Systems   •   85

Then, the length of the cable is given by the following integral. Note
we are reflecting the parabola across the entire y axis to complete the pa-
rabola, which doubles the area—hence the 2.

2
1,400  ft  2  1,400  ft 4
Length = 2 ∫ 1+  x    dx = 2 ∫ 1+ x 2   dx
0
 7,000  0 4.9  × 10 7

For simplicity, let’s call the constant factor on the x squared term,
1/C2 (yes, the relabeling seems like overkill, but the swap will make the
analysis a bit cleaner). If we factor out 1/C2 from under the radical, then
our integral becomes

1,400  ft 1 2 2 1,400  ft
Length = 2 ∫ 1+ 2
x   dx = ∫ C 2 + x 2   dx
0 C C 0

Now, the usual integration tricks will not work on this integral. In-
stead, we need to use a technique typically referred to as “trig substitu-
tion.” Our goal here is not to demonstrate the integration technique, rather
we want to explore the subtle differences in models we use for the same
objects and phenomena. Therefore, we will spoil the surprise and make the
following substitutions:

x = C tan (θ ) ,   dx = C sec 2 (θ )  dθ

Note we also must change the bounds on the integral when we


e­ xchange variables. Our new variable is θ , so we need to know what the
value of θ is given the value of x. For the lower bound, we set x = 0 and
solve for θ

0 = C tan (θ )

θ = tan −1 ( 0 ) = 0

And for the upper bound, we set x = 1, 400 and solve for θ once more:

1, 400 = C tan (θ )

 1, 400 
θ = tan −1  ≈ 1.57
 C 
86  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Making all of the substitutions results in:

2 1.57
C 2 + C 2 tan 2 (θ )  sec 2 (θ )  dθ
C ∫0
Length =

The C 2 term is becoming quite a pest, isn’t it? We can deal with
it again. Factoring it out of the radical and the integral not only causes
­cancelation with the 1/C term outside the integral, but also reveals the
trigonometric identity, 1 + tan 2 (θ ) —which can be simplified to sec 2 (θ ) .
Performing the previous actions results in

1.57 1.57
Length = 2 ∫ 1 + tan 2 (θ )  sec 2 (θ )  dθ = 2 ∫ sec 2 (θ )  sec 2 (θ )  dθ
0 0
1.57
= 2∫ sec 3 (θ )  dθ  
0

Amazingly, our integral simplifies to a deceptively simple problem.


The integral of sec 3 (θ ) is a beast of a problem, so much so it has its own
Wikipedia page.3 Talk about mathematic fame! Well, infamy.
Let’s look at the other model instead. At first glance, it looks more
complicated than the parabolic model:

  x  
y2 ( x ) = 3, 545.708  cosh   − 1
  3, 545.708  

Taking the derivative is easy, if we realize that the hyperbolic cosine’s


derivative is hyperbolic sine:

 x 
y2 ′( x ) = sinh 
 3,545.708 

Next, we plug y2 ′( x ) into our formula:

2
1,400  ft   x 
Length = 2 ∫ 1 + sinh      dx
0
  3,545.708 

Notice this integral will yield to a simple integration technique,


u-substitution. We can assign a new variable (u) in place of a complex

3
“Integral of secant cubed.” https://en.wikipedia.org/wiki/Integral_of_secant_cubed
Analyzing Failure in Systems   •   87

quantity. Using u will vastly simplify our expression. Importantly, the


bounds will change due to the exchange of variables:

x 1
u= ,   du =   dx
3, 545.708 3, 545.708

1 + [ sinh ( u )]   du
0.7897 2
Length = 2(3,545.708) ∫0

Like before, we just uncovered another identity, 1 + sinh 2 u —which


is equivalent to cosh 2 u . If we make that replacement, we have

0.3948 0.3948
Length = 7,091.416 ∫ cosh 2 u   dx = 7,091.416 ∫ cosh( u )  du
0 0

Compared to our previous model, finding the length of the cable is


trivial here—even though our formulation was somewhat more involved.
We have the option of picking between y1 ( x ) and y2 ( x ) , our choice
depends on what our goal is. In this case, we wanted to find the length
of the cable—by hand, no less. Clearly y2 (t ) provided the path of least
resistance, but we could do so because the difference between y1 ( x ) and
y2 ( x ) for small values of x (less than 3,000) appears to be negligible based
on Figure 4.1. Since we know our answer will be approximately the same
as our previous model, we can find a rough estimate by hand—­revealing
that the cable was approximately 3,100 ft in length.

 0.3948 
7,091.416 ∫
0.3948 
cosh( u )  du = 7,091.416 sinh ( u )  ≈ 2,873 ft 
0 
 0 

What’s the moral of this convoluted journey? We can have multiple


models for the same phenomena, but one model may provide advantages
over others either in analysis or in realism. Each model has its purpose,
and it is up to us to evaluate the benefits and costs associated with each
of our models.
***
With the length known, we’d like to check whether the main cable is
going to fail under the wind conditions. To do so, we need to know the
internal force, F. Let’s start with the facts: the cables themselves were
88  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

about 15,625 kg per cable with a diameter of approximately 1 in. using our
results from the previous calculation. Based on observations, the torsional
rotation cycled every 4.3 seconds with an amplitude of 4.2 m, which is the
displacement of the bridge deck from its initial position.4
That’s the solid evidence, but to keep going, we need to name some
assumptions. Let’s assume that the load is applied halfway through the
oscillation of the deck. Then, we can say that the duration of that cycle was
1.1 seconds. From these assumptions and the straight facts, we can use the
following equation to find the force in the cables:

2 g∆x
F =m
t

where m is the mass of the cable, g is gravitational acceleration, ∆x is the


change in bridge deck’s starting position, and t is the duration of the cycle.
If we calculated the internal force in the cables, we would obtain

 m
2  9.81 2  ( 4.2   m )
 s 
F = (15, 625  kg ) = 129   kN
1.1 s

using the mass of the cables, the amplitude of the bridge’s displacement,
and the duration of the cycle.
We now know the internal force, which means we can start looking
at the engineering stress in the cables. Sounds intimidating, but it’s quite
a simple formula. The engineering stress is just a measure of the inter-
nals forces that the particles the material is composed of exert on each
other. The stress can be found by taking the force and dividing it by the
cross-sectional area:

Cross-sectional area

Idealized cable

4
B. Fillenwarth. 2007. “Improving the Mathematical Model of the Tacoma Narrows Bridge.”
Rose-Hulman Undergraduate Mathematics Journal 8, no. 2, p. 7.
Analyzing Failure in Systems   •   89

If we calculate the internal stress by taking the force we just calcu-


lated and divide it by the cross-sectional area of the cable, which is the
radius (1/2 in. or 0.0127 m) squared times pi,

129   kN
σ = 2 = 254.6   MPa  ( megapascals )
π ( 0.0127   m )

If we compare the stress to the yield stress of the cables, which is


between 250 MPa and 350 MPa, we will note that the internal forces are
barely high enough to cause the cables to yield—that is, begin to deform.
This would imply, based on our simple model, that the cables failing and
resonance are not sufficient explanations for the collapse. In fact, the use
of the Tacoma Narrows Bridge as an example of harmonic motion and res-
onance has been criticized as an oversimplified explanation for the failure,
so our result is not too surprising.5

4.2  THE CASE OF THE I-35 WEST MISSISSIPPI


BRIDGE

An eight-lane, steel truss arch bridge—officially known as Bridge 9340—


used to exist in Minneapolis, Minnesota. During rush hour in the evening
of August 1, 2007, the bridge collapsed without warning. In the wake of
the bridge failure, 13 people were killed and 145 were injured. On the day
of the collapse, four of the lanes were closed due to construction, with
roughly 575,000 lb. of supplies idle in the restricted lanes—this circum-
stance was among the first causes to be identified.
Leading up to 2007, the bridge was no stranger to poor ratings from
the government and other reviewing bodies, such as universities. In
2001,  the  civil engineering department at the University of Minnesota
evaluated the bridge. During the inspection, cracks were found in the
cross-girders that connected to the main spans of the bridge. The depart-
ment made recommendations to implement strain gauges and continue to
monitor the structural integrity of the bridge.
Later in 2005, the bridge was rated as “structurally deficient” yet again
with a replacement being the likely solution (www.infrastructurereportcard
.org/cat-item/bridges). Despite numerous warnings, the governor at the
time chose to save that problem for later and announced that the bridge
would be completely replaced starting in 2020. We may think, “Well, that’s

5
B.Y. Billah, and R.H. Scanlan. 1991. Resonance, Tacoma Narrows bridge failure, and
­undergraduate physics textbooks. American Journal of Physics 59, no. 2, pp. 118–124.
90  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

not too bad. It could have lasted until then.” To put into perspective how
poorly this bridge performed, out of 100,000 heavy-traffic bridges, only
4 percent were rated below it.

ASCE’S INFRASTRUCTURE REPORT CARD

We have seen two examples of failing bridges, but these must be


anomalies, right? Collapsing bridges are indeed rare; however,
structurally deficient bridges are actually more common than you
might expect. According to the American Society of Civil Engineers
(ASCE), 9.1 percent of the nation’s 614,000 bridges are rated as
structurally deficient. This means that, in 2016, 188  million trips
were taken over 56,000 structurally deficient bridges each day.
The ASCE gives this a grade of C+ on their Infrastructure Report
Card. Other grades include a D for drinking water: while Americans
have relatively easy access to clean drinking water, treated drinking
water is easily lost: 6 billion gallons/day through leaking pipes and
240,000 water main breaks are a result of the $1 trillion needed to
repair and expand systems to meet needs.6

To determine the cause of the failure, a report by the National Transpor-


tation Safety Board (NTSB)7 suggested that the gusset plates failed due to two
main reasons: increased weight of the bridge from modifications from the last
few years and the traffic/construction loads. Let’s look at the gusset plate it-
self, the U10 plate. We can start by checking the forces acting on the U10 plate
by drawing a free body diagram and using principles of static equilibrium.

DETERMINING THE REQUIREMENTS FOR EQUILIBRIUM


IN THE U10 PLATE

We’re going to use this free body diagram to see what it takes to keep this
plate in equilibrium at the critical sections (Figure 4.2). Our job is to de-
termine the important forces on top: the equilibrating shear (V) and axial
(P) and the moment (M). Once an engineer determines the forces that keep
the plate stationary, they can begin to compare stresses and limitations

6
Infrastructure Report Card. 2017. “Making the Grade.” Infrastructure Report Card. https://
www.infrastructurereportcard.org/making-the-grade
7
National Transportation Safety Board report on the Failure of the I-35 Highway Bridge.
https://www.ntsb.gov/investigations/AccidentReports/Pages/HAR0803.aspx
Analyzing Failure in Systems   •   91

14 in.

100 in. (width)


72 in. (height)
F3
F1
F2
A C
B
63.87 55.05
53.33
39.83

38.0 38.0

A B C

Figure 4.2.  Diagram of the U10 plate and internal forces.

in order to determine its overall performance—which will clue us in to


exactly how this plate failed. We know the forces in the beams themselves
(F1, F2, and F3), so those are given to us through field measurements8:

F1 = 2, 288  kips

F2 = 540  kips

F3 = 1, 975 kips

8
R. Holt, & J.L. Hartmann. 2008. Adequacy of the U10 & L11 gusset plate designs for the
Minnesota Bridge No. 9340 (I-35W over the Mississippi River). Federal Highway Adminis-
tration, Turner-Fairbank Highway Research Center.
92  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Figure 4.3.  Free body diagram of the U10 gusset


plate.

Note that the unit here is kip, otherwise known as 1,000 lb of force. The
conversion to Newtons is the following: 1 kip = 4,448.2 Newtons (N) =
4.4482 Kilonewtons (kN). First, let’s redraw the free body diagram without
some of the visual clutter (Figure 4.3).
All we need to do is use the most fundamentals principles in Statics,
a special case of Newton’s Second Law where the sum of all forces and
moments is zero.

∑F = 0    and   ∑ M = 0   

Notice there are two right triangles at the end of the beams. The trian-
gles are not there for decoration or to complicate the picture; rather, they
handle all the trigonometry without any utterance of sine or cosine. You
may see that the dotted lines feed into the hypotenuse of the triangle, which
means the angle formed by that triangle is the same as the angle of the force.
These lifesavers allow us to easily do trigonometry and calculate the values
we need to break the forces into components. To illustrate, the famous 3-4-5
triangle and some force, F, with an angle, θ is shown in Figure 4.4:
F Our job is to break F into its
Fy
components, Fx and Fy  . If we started
5
4 with Fx then our relation will need to
Fx include the base of the triangle. That
3 means the trig function we’d want to
use is cosine—adjacent (base) over
Figure 4.4.  3-4-5 Triangle. hypotenuse (Figure 4.5).
Analyzing Failure in Systems   •   93

F adjacent
cos ( ) =
hypotenuse

5 3
4 cos( ) =
Fx 5
3

Figure 4.5.  Computing Cosine from 3-4-5


Triangle.

Now, if we extended this to the force itself, then we just need to do the
trig on the big triangle. The adjacent side will be the x part of the force, Fx ,
and the hypotenuse will be the whole force, F.

Fx
cos (θ ) =
F

We can work some algebraic magic to solve for Fx. Multiply both
sides by F.

Fcos (θ ) = Fx

But wait a minute, we know the cosine of that angle—it’s 3/5! We


can swap cos(θ ) for 3/5 in our equation and arrive at a result that’s easy
to calculate! This means we can think of the triangle as a tool to scale our
total force to match its components in the x and y directions.

3
Fx = F
5

Solving for the x-components: Using the reference triangle like we did
with the 3-4-5 triangle, we know

opp 38.0
cos (θ ) = =
hyp 63.87

If we use the formula from before, then we can write the x-component
of the first force like so

 38.0 
F1 x = F1 cos (θ ) = ( 2, 288 ) 
 63.87 
94  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

This means that F1 x = 1, 361.265 kips. Since F2 is vertical, it will


have no x-component. Now, we apply the same idea for the x-component
for the force, F3, which should have the form F3 x = F3cos (θ ) . As before,
we just need to use the cosine relation:

opp 38.0
cos (θ ) = =
adj 55.05

Then we take that “cosine factor” and apply that to the trig of the
bigger triangle:

 38.0 
F3 x = F3 cos (θ ) = (1, 975) 
 55.05 

F3 x = 1, 363.306 kips

We start by using a fundamental principle in Statics—the sum of the


forces in the x direction is equal to zero. Using this fact, we can determine
the magnitude of the equilibrium shear, V. We usually pick the positive
direction to be vectors pointing to the right.
Looking at our free body diagram (Figure 4.6), the x-components
of the forces we know both point in the positive direction; however, the

Figure 4.6.  Finding the relevant forces to sum in the x direction


(in gray).
Analyzing Failure in Systems   •   95

shear is pointing in the negative direction. That means the following


is true:

F1 x + F3 x − V = 0

V = F1 x + F3 x

We already know what those two forces are equal to, so we can sub-
stitute those values in

V = (1, 361.265) + (1, 361.265)

V = 2, 722.53 kips

Next, we will find the y-components and use them to determine the
axial force P.
We apply basic trigonometry and similar triangles again. This time we
use the sine of the angle:

adj 51.33
sin (θ ) = =
hyp 63.87

Applying that idea to the component itself

 51.33 
F1 y = F1 sin (θ ) = ( 2, 288 ) 
 63.87 

We then find that F1 y = 1, 838.783 kips. Now we need to find the


y-­component of the third force.

adj 39.83
sin (θ ) = =
hyp 55.05

Then we take that “cosine factor” and apply that to the trig of the
bigger triangle:

 39.83 
F3 x = F3 cos (θ ) = (1, 975) 
 55.05 

F3 y = 1, 428.960 kips
96  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Finding the axial force, P—To determine the axial force, we need to
use the same principle of Statics as before. This time, we know the sum
of the forces in the y direction is zero. We choose forces pointing up to be
positive:

Figure 4.7.  Finding the relevant forces to sum in the y


direction (in gray).

Looking at the free body diagram, we can see that F1 and P are point-
ing in the positive direction while F2 and F3 are negative (Figure 4.7).
Writing that as an equation,

F1 y + P − F2 − F3 y = 0

Rearranging the equation to solve for P,

P = F2 + F3 y − F1 y

Now we can substitute in those values:

P = (540) + (1, 428.960  ) − (1, 838.783)

P = 130.177 kips
Analyzing Failure in Systems   •   97

The equilibrium moment is the last


piece of information we need to find, F
the rotational potential that keeps the d
object from rotating. Moments are cal-
culated a bit differently; for a moment
A
around some point, we just need to
know the magnitude of the force and its Figure 4.8.  Demonstration of how
perpendicular distance from the point. a moment is calculated.
To find the value of the moment around
a point (like A in the middle), we just multiply the force by its distance
from A, the simplest model is shown in Figure 4.8.
Just like determining forces, choosing a positive direction is im-
perative to finding the correct value. To calculate the moment about the
crossed point above the plate, we need to use the idea that the sum of the
moments is equal to zero—we consider counterclockwise to be positive.


+ ∑M =0

Moments are the tendency of the object to rotate due to some applied
force. That means we need to consider how far our forces are from the
point we’re considering—which would be 14 in. Recall the equation to
determine moments is given as

M = Fd

where F is the applied force and d is the distance from the considered
point. All the forces are 14 in. away and the x-components will produce
motion. What about the y-components? Think about the plate as a ball
attached to a steel rod from the crossed point. At the moment (no pun
intended), nothing is moving, and we want the forces to produce motion.
Will pulling the ball attached to the steel rod up or down cause it to move?
Unless we rip the rod and ball from the wall with our incredible strength,
y forces are certainly not going to budge the ball—just like they are not
going to budge the plate given our reference point. With that in mind, let’s
write the equation:

F1 x d + F3 x d + M = 0

Rearrange the equation to solve for M:

M = − F1 x d − F3 x d
98  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Now let’s substitute in the values we know. Both forces are 14 in. away,

M = −(1, 361.265)(14 ) − (1, 363.306 )(14 )

M = −38,144   kips − in.

Since the moment is negative, the direction the plate will tend to turn
would be clockwise.

M = 38,144 kips − in.   clockwise

Now that we know the forces keeping the plate in place, we can talk
more specifically about the material it was made of and how the material
and forces are related.

MATERIAL PROPERTIES OF THE U10 PLATE

By referencing the official documentation on the I-35W bridge, we can


determine exactly what material the members feeding into the gusset plate
were made of and their properties. Now that we know the forces, all we
need to know is how the material behaves on the basic level of microstruc-
ture (what we can’t see with our own eyes). From the documentation, we
can infer the beam is composed of grade 50 mild steel.

Young Modulus, E

While the official title of this material property doesn’t tell you much
about its definition, its alternate name is the elastic modulus. Elasticity—
that probably made you think of elastic bands (or rubber bands). Naturally,
the elastic modulus is a measure of the stiffness—its resistance to deform
from some applied force—with the common unit of gigapascal (GPa).
This means that the higher the material’s modulus of elasticity, the more
stiff the material, and the less likely it will deform from heavy loads. We
can easily calculate the elastic modulus using the tensile stress (internal
forces of particles on each other) and extensional strain (how much the
material deforms). The symbol for tensile stress is commonly sigma, σ ,
while epsilon, ε , is used for extensional strain.

σ
E =
ε
Analyzing Failure in Systems   •   99

Steel typically has an elastic modulus of 200 GPa, which makes it


fairly stiff.

Poisson’s Ratio, v

This quantity is a little bit more complicated to explain. Poisson’s ratio


is a dimensionless measure of how a material expands in the two direc-
tions perpendicular to the applied force and parallel to the flow when
compressed. Maybe a picture can make this easier—below is a cylinder
being compressed by two forces of the same magnitude, F. The ratio is
formed between the latitudinal strain (width-wise, horizontal) represented
by ε lat and the longitudinal strain (height-wise, vertical) represented by
ε long (Figure 4.9).
Poisson’s ratio is given the Greek letter nu (v).

ε lat
ν =−
ε long

Due to the properties of the elastic modulus, Poisson’s ratio can only
be between −1 and ½, theoretically. The reality is, most common materi-
als usually fall between 0 and ½:

1
0≤ν ≤
2

The steel used in the I-35W bridge has the property, ν = 0.3. What
does this mean for our material? A high Poisson’s ratio will indicate that
a material will become narrower as it is pulled apart (think of a rubber
band as it’s stretched). A value near zero means that a material will stretch
without changing much in the transverse direction—it will keep stretching
like silly putty. When Poisson’s ratio is negative, the material will tend to
stretch in all directions.

F
long

lat lat

long
F

Figure 4.9.  Illustration of Poisson’s ratio.


100  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Yield Strength and Ultimate Strength

The last two qualities we’ll be looking at are concerned with the behavior
of the material under loading. These can be picked off of a stress–strain
curve—a graph that shows us the points where a material will reach its
critical points. Stress–strain curves are already developed through exper-
iments in a laboratory and available for engineers to reference. The yield
strength, σ Y , is the point where a material essentially gives up trying to
­resist the force and begins to deform plastically (permanently). The ulti-
mate strength, σ u , on the other hand is the maximum force that the material
can possibly handle, which means the material will start to fail. Both the
yield and ultimate strength are measured in megapascal (MPa) (1 lb/in.2 =
0.006895 MPa).
Finally, we reach fracture, where the material fails. Figure 4.10 is
the stress–strain curve for the mild steel in the bridge, with the three main
pieces identified on the curve in traffic light sequence.
A particular type of fracture is failure by shear. This occurs when
enough force is applied in order for the internal surfaces to slide past one
another. Take the block below for an idealized example of some material
failing by shear (Figure 4.11).

Yellow Light – Ultimate Strength


Material reaches the highest amount
of force it can handle, begins to fail.

Red Light – Fracture Material


fails (breaks, snaps, etc.)
Green Light – Yield Strength
Stress

Material begins to deform.

Strain

Figure 4.10.  Stress–strain curve for a metal.

Figure 4.11.  Illustration of shear.


Analyzing Failure in Systems   •   101

Table 4.1.  Shear strength

Ultimate shear Yield shear


Material strength, σ SU strength, σ SY
Steel 0.6 × σ u 0.58 × σ Y
Ductile iron 0.90 × σ u 0.75 × σ Y
Aluminum and alloys 0.65 × σ u 0.55 × σ Y

The material susceptibility to shear is closely related to the yield and


ultimate strength. You’ll often find that a material’s ultimate shear strength
is much lower than its typical strength (Table 4.1).
When the mild steel used in the bridge is tested, the material began
to deform at 348 MPa and began to fail at 593 MPa. Based on the stress–
strain diagram, the value where the material begins to succumb to the
force is the yield strength and the point at which the material reaches the
highest amount of force and begins to fail is the ultimate strength. There-
fore, σ Y = 348 MPa and σ u = 593 MPa. In addition, the steel used in
the plate is said to have a Poisson’s ratio of 0.3. Since v can only be
­between −1 and 1/2, we can consider that it’s closer to the 0.5 descrip-
tion. This means that the steel is more likely to become narrower as it is
pulled apart.

THE STRENGTH OF MATERIALS: THE U10 PLATE

Knowing the material properties, we’re able to determine what’s going


to happen to the U10 gusset plate when it’s subjected to shear. We’ll be
mainly concerned with the shear stress, a force that acts tangential to the
material’s surface. Think about a nail that goes through a pair of 2×4’s.
Instead of pulling the board straight out, say we pulled straight up or down.
This action would result in the nail being subjected to simple shear since
the force is attempting to cut through the material. Take a look at the bolt
below, there are two loads (the rectangular-shaped boxes with arrows) act-
ing it. This could be due to something pushing down on the bolt and the
material it’s lodged in resisting the force (Figure 4.12).
In the case of the gusset plate, the metal is ½ in. thick and is connected
to the steel members by rivets. In the original design, the gusset plate was
intended to be 1 in. thick, but design changes cut the overall thickness
in half.
102  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Figure 4.12.  Forces acting upon a bolt in opposite


directions.

Below, we have our gusset plate with two forces acting in opposite
directions in such a way that the gusset plate is subject to shear. The Amer-
ican Association of State Highway Officials (AASHO) only allows 15,000
lb/in.2 (about 103.4 MPa) of stress—exceeding this is considered a safety
hazard. Now, we need to check and see whether our gusset plate satisfies
that condition. Assuming the upper section is approximately 100 in. with
a thickness of ½ in., we can use the following equation to determine the
shear stress (Figure 4.13):

V shearing force
σV = =
A cross − sectional area

V1 100 in. V2

VR

Figure 4.13.  Shear on the U10 plate.

In the situation above, the loads were estimated to be V1 = 949  kipf


and V2 = 2,147  kipf .

a. Determining the resultant shear—This is just vector addition. Both


V1 and V2 are vectors with x-components only, so we can just add
those two vectors together.

VR = V1 + V2

VR = ( 949   kips ) + ( 2,147   kips ) = 3, 096   kips


Analyzing Failure in Systems   •   103

b. If we want to calculate the shear in the upper half of the gus-


set plate, then we just need to know the force acting tangent to
the plate and the cross-sectional area. We know the upper half is
100 in. across and ½ in. thick, so the cross-section is the shape of
a rectangle.

1
Cross − sectional area = bh = (100   in.)( in.) =   50   in.2
2

We know the resultant shearing force, so now we can just plug those
into the formula:

V 3, 096  kips kips


σV = = 2
= 61.92 2
A 50  in. in.

c. AASHO Comparison—Compared to the AASHO requirements,


anything over 15,000 lb/in.2 or about 103.4 MPa is not considered
appropriate. Our units don’t quite match up, but it should be clear
that the shear stress we found violates the AASHO code.

kips lb
61.92 2
> 15, 000 2
in. in.

If you’re not convinced, remember that 1 kip is the same as 1,000 lb,
so 61.92 kips/in.2 is just 61,920 lb/in.2. So clearly,

lb lb
61, 920   > 15, 000 2
in.2 in.

Will it shear?—To answer this question, we need to refer back to the


“Material Science” section of the scenario. We’re looking for the ultimate
shear strength, which is given in the short chart to be

σ SU = 0.6   ×  σ u

for steel, which is what the gusset plate is made of. We know the ultimate
strength of the steel is 593 MPa, but we need to adjust that for the ultimate
shear strength:

σ SU = 0.6   ×   ( 593 MPa ) = 355.8  MPa


104  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Now, if our shearing force exceeds the ultimate shear strength, the
plate will shear:

If   σ V > σ SU ,   then the plate will shear

Let’s convert our units to MPa:

lb 0.006895 MPa
61,920 2
× = 426.94 MPa
in. 1 lb/in.2

Now we can directly compare the two stresses:

σ V > σ SU

( 426.94   MPa ) > ( 355.8  MPa )

The shearing force is much larger than the ultimate shear strength,
which means: yes, the plate will shear.

d. Calculating the shear stress—This time we need to calculate the


shear stress assuming the original design was used. That means we
need to recalculate the cross-sectional area:

Cross − sectional area = bh = (100   in.)(1  in.) =  100   in.2

Then we can plug our result into the formula for shear stress:

V 3, 096   kips kips


(σV )1  in.   = = = 30.96 2
A 100   in.2 in.

Comparing stresses—Let’s do a quick conversion to get our newly


calculated stress to get it in MPa like the others.

lb 0.006895 MPa
30, 960 2
× = 213.5 MPa
in. 1 lb / in.2

Now, will the plate shear? Let’s use the same inequality as before.

(σV )1  in.  > σ SU

( 213.5   MPa ) < ( 355.8 MPa )


Analyzing Failure in Systems   •   105

It turns out the shear stress is less than the ultimate shear stress, so
the plate isn’t going to shear! In fact, you may notice that doubling the
cross-sectional area nearly cut our shear stress in half!

1
(σV )1  in. ≈ (σV )1/ 2   in.
2

The new stress is still in violation of the AASHO recommended value,


but is certainly safer than what was implemented.

STRUCTURAL LOADS: GOING BEYOND THE BRIDGE

But our situation here is simplified, all the loads on the bridge were not
concentrated at a single point, rather they were distributed over a given
area like the loading w ( x ) in Figure 4.14.
A structural load is a force that is applied to some structure or its
­components—a load can be as simple as a truck (the load) on a bridge (the
structure). These forces can tend to cause the structure to deform and even
fail depending on the load. Most loads aren’t going to be concentrated at
one point. Just think about laying on a memory foam mattress, your weight
is going to be spread out across your body—which would be an example
of a distributed load. These forces are applied over a certain distance that
can either stay constant or vary. A continuous distributed load can be de-
scribed as a function, we’ll call it w ( x ). In Figure 4.12, there’s a continu-
ous distributed load that’s pressing down on the beam over some distance.
We can replace that nasty load with a single force using an integral. In fact,
the total force is just a definite integral. The bounds are just the starting
point, di, to the final point, df.
df
W = ∫d w ( x )  dx
i

w(x)

di df

Figure 4.14.  An example of a distributed load.


106  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

That’s good and all, but where does that resulting force go? Does the
term centroid sound familiar? If we recall the term from geometry, the
centroid is the official geometric center of a shape where it will balance
perfectly on a needle. The centroid is just a point in space, so it is given as
coordinates, C = (Cx  , C y ), for a two-dimensional object. Simple enough
until we consider what we need to do to find out what the centroid is for
a shape. In fact, given the shape traced out by the function f ( x ) over the
interval a to b, the coordinates of the shape’s centroid turn out to be fol-
lowing integrals:

Cx =
∫a xf ( x )  dx
b
∫ a f ( x )  dx
b
∫ a  f ( x )   dx
2

Cy = b
2 ∫ f ( x )  dx
a

Example 4.1: Finding the centroid of basic shapes (taking advantage


of functions)

Let’s find the centroid of a triangle by using the fact we can make a
triangle using the function f ( x ) = x . To make the calculations rela-
tively easy, let’s make a triangle of side lengths 1 on the interval [0,1]
(Figure 4.15).

Figure 4.15.  Triangle created using the


function f ( x ) = x on the interval [0,1].
Analyzing Failure in Systems   •   107

With most problems, half the time we spend will be in setting up the
solution. For the x-component, we use the formula

Cx =
∫ axf ( x )  dx
b
∫ a f ( x )  dx
We’re using f ( x ) = x on the interval [0,1], so our formulas become

b 1
∫ axf ( x )  dx =
∫ 0x( x )  dx
b 1
∫ a f ( x )  dx ∫ 0x   dx
Both integrals can be done easily, and they both use power rule. This
means the x-component of the centroid is determined by

Cx =
∫ 0x 2   dx
1
∫ 0x   dx
Once we integrate, we still need to figure out what the total area is
going to be by evaluating each integral at the end points:

1 3 1
1 x
Cx =
∫ 0x 2   dx =
3 0
1
∫ 0x   dx 1 2 1
x
2 0

By substituting x = 1 into the numerator and denominator ( x = 0 will


not add anything to the area), we find the x-component of the centroid is 2/3:

1
Cx = 3 = 2
1 3
2

For the y-component, we need to use a slightly different formula:

b 1
∫ a  f ( x )   dx ∫ 0 [ x ]   dx
2 2

Cy = b = 1
2 ∫ f ( x )  dx 2 ∫ x   dx
a 0
108  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Those integrals look awfully familiar, don’t they? We don’t need to


waste our time redoing them.
1

Cy =
∫ 0x 2   dx =
1
3
1
2 ∫ x   dx
0
( )
2 12

After simplifying, we see that the y-component is 1/3.

1
Cy =
3

Put the two together, and we’ll get our centroid!

 2 1
(
C = Cx  , C y =   , 
 3 3 )

Figure 4.16 shows us where the point C would be.

Figure 4.16.  Placement of the centroid.

It should be no surprise that the centroid is going to be the place the


resulting force needs to be applied. The resulting force is the concentration of
the load much like the center of mass is the concentration of most mass. Luck-
ily, the formulas remain unchanged mathematically. It’s only been relabeled to
keep consistent with our notation here. If you look back to when we calculated
the centroid for shapes, you’ll see that these are the same formulas at heart.

df
∫ d xw ( x )  dx
C = d
i

∫ d w( x )  dx
f

i
Analyzing Failure in Systems   •   109

Table 4.2.  Two simple special cases of distributed loads

L
L

d
d

Rectangular Load Triangular Load


Whenever the load, L, is constant If the load, L, forms a right trian-
over a distance, d, we can just use gle over the distance, d, then we
the area formula for a rectangular can use our usual right triangle
we’ve used forever. The resultant formula. The resultant force W is:
force W is:
1 1
W = Ld = ( height )( base )
W = Ld = ( height )( base ) 2 2

For the centroid, it’s going to be The placement of the centroid can
in the center (half of the distance) be found in two different ways. It
since the load is constant. can be 1/3 of d from the leg or 2/3
of d from where the triangle ends.
1
Cx = d
2

1 2
/3 of d /3 of d

Luckily, we can use some tricks to skip the integration process. There are
two special cases that are useful to highlight—rectangular and triangular
loads (Table 4.2).
***

Example 4.2: Replacing a distributed load with a single force:


Geometric method

To demonstrate how to simplify a distributed load into a single force, let’s


use the geometric shortcuts first, then try the general approach using the
integral (Figure 4.17).
Let’s set up the integral like we do not know w ( x ) . Since the load
starts at 2 ft and ends at the 4 ft mark, those will be our bounds.

df 4
W = ∫d w ( x )  dx = ∫2 w( x )  dx
i
110  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Figure 4.17.  A distributed load on a 6 ft long beam.

Imagine this, we don’t even need to know a formula for w ( x ) or even


do an integral. The area is a rectangle, with a base of 2 and a height of
500 (lb). Since we know the formula for the area of a rectangle, we can
solve for the magnitude of the resultant force:

4
W = ∫2 w( x )  dx = ( 2 )(500) = 1,000 lb

Now for the centroid, C x . If the load was not such a simple shape, we
would need to use the integral formula. In this case, we have a r­ ectangle—
meaning the load is even distributed over a given distance. Due to the
uniform distribution, we know that the centroid of the rectangle is right in
the middle of the affected area:

df
∫ d xw ( x )  dx
Cx = i
d = 3  ft
∫ d w ( x )  dx
f

Now that we know the resultant force and its position on the beam, we
know the rectangular distributed load is equivalent to a single force with
a magnitude of 1,000 lb centered at 3 ft from the left end of the beam—as
shown in Figure 4.18.
We should note that w ( x ) can be broken down into simpler geometric
shapes (see Figure 4.19). Just like integrals, the areas can be calculated
separately and then added together to find the grand area. In the context
of distributed loadings, each piece of the load will have its own resultant
force, which combine to form a resultant load.
Analyzing Failure in Systems   •   111

Figure 4.18.  Resolution of a distributed load into


a single force.

w1 (x)

w2 (x)

Figure 4.19.  Breaking a load into pieces to simplify calculations.

***

Example 4.3: Replacing a distributed load with a single force:


Integral method

Next, we will resolve a nonlinear distributed load into a single force. Here
we will describe the distributed load as w ( x ) = 5 x —clearly not a rectan-
gle or triangle—as pictured in Figure 4.20.
We will start by setting up the integral. We have a function, w ( x ),
to describe the load in kilonewton, so that becomes the integrand. The load
starts at 0 m and ends at 4 m, which will become the bounds on the integral:

df 4
W = ∫d w ( x )  dx = ∫0 5 x   dx
i

Let’s rewrite the radical using exponents to make it more “integral


friendly,” then apply the power rule:

4
4 5 6/5
W = ∫0 x1/ 5   dx = 6
x  
0
112  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Figure 4.20.  Nonlinear distributed load on a 7 m beam.

We use our basic calculus principles, the Fundamental Theorem of


Calculus, to find out the resultant force is 4.4 kN.

4
5 6/5 5 5
W = x   = ( 4 )6 / 5 − (0)6 / 5 = 4.4  kN
6 6 6
0

How about the centroid? Since we have a nonstandard shape for the
distributed load, we will need to use the centroid formula. We just need to
calculate one more integral since we just calculated the total area in the
denominator (in gray):

∫ 0x ( x1/ 5 )  dx
df 4
∫ d xw ( x )  dx
Cx = i
d = 4
∫ d w ( x )  dx ∫ 0x1/ 5   dx  
f

The denominator becomes 4.4 and the integral on the numerator is a


simple application of the power rule like the previous integral:

5 11/ 5 4
4 x
Cx =
∫0 x 6 / 5   dx
=
11 0
4.4   4.4  

By plugging x = 4 into our antiderivative, we find the numerator is


equal to 9.6, which then becomes 2.18 m when divided by the total area of 4.4.

9.6  
Cx = = 2.18  m
4.4  
Analyzing Failure in Systems   •   113

Figure 4.21.  Resolution of the nonlinear distributed load


into a single force.

Using our results, we can redraw our original distributed load as a


single force with a magnitude of 4.4 kN that is 2.18 m from the left-hand
side of the beam (Figure 4.21).
***

FINITE ELEMENT ANALYSIS OF THE U10 PLATE

In the past, we needed incredible mathematical prowess to find solutions


to some difficult problems. Before Calculus was discovered, dealing with
infinitely small quantities was next to impossible. In fact, often the solu-
tion cannot be expressed using things we know, even today—meaning they
must be solved numerically. With that in mind, we can’t directly dive head
first into the problem, we must solve it indirectly using an algorithm to
approximate the solution. Approximation is the key word here, we’re not
going to get a neat and clean answer that can be spoon-fed to us. This is
the answer, in the rawest form, derived using the numerical technique of
finite element analysis.
In a sense, we’ve had a taste of the goals of finite element analysis
(FEM) when we looked at approximating area under the curve in Calcu-
lus. The definite integral provides the solution to the continuous problem,
which essentially means we were able to find the answer to the problem
in terms of mathematical objects we can easily describe. Take this integral
for example:

8 x4 + x7
∫0 ( x + 5)2   dx

There is little we can do here; the antiderivative of this function is


complex to calculate by hand. It makes more sense to use numerical
114  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

integration to find the solution instead of integrating the old-fashioned


way. Remember where the definite integral comes from? It was this limit:

 n 
lim  ∑ f ( xi* ) ∆x  =   ∫ f ( x )  dx
b

n→ ∞ a
 i =1 

This idea came from fitting a bunch of rectangles under the curve
to approximate the area. In theory, if we keep increasing the number of
rectangles, the closer we’ll be to finding the true area under the curve.
Sectioning off bits and pieces of the area and calculating them on their
own is a kind of discretization. By discretizating our continuous problem
that is fairly challenging to solve, we can transform the hard problem into
a series of much easier problems. The solution to the hard problem is then
the sum of the easier problems (Figure 4.22).

f )


8
f )
( )
f )

Figure 4.22.  Approximation of the integral


using three trapezoids (elements).

The idea with numerical integration is to use an approximation tech-


nique to estimate the area and avoid integration. In the picture, we’re using
the trapezoidal rule. Unlike the method of using Riemann sums, we are
taking the intervals and approximating them using a trapezoid rather than a
rectangle. In GeoGebra, an easy-to-use dynamic geometry software appli-
cation, this function is easily found. All we need to do is type the following:

TrapezoidalSum[<function>, <start x-value>, <end x-value>,


<number of trapezoids>]

This means that our picture above can be recreated by defining the
function f and graphing it:

f(x) = (x⁴ + x⁷)^(1/2)/(x + 5)²


Analyzing Failure in Systems   •   115

Table 4.3.  Comparison of convergence to the area under the curve for
different sums

Lower Upper
Number of Riemann Riemann Trapezoid
Step elements sums sum rule
1 3 10.21976 33.09263 21.65620
2 10 16.90604 23.76790 20.33697
3 100 19.86566 20.55185 20.20876
4 1,000 20.17316 20.24178 20.20747
5 1,0000 20.20403 20.21089 20.20746

Then we can get the trapezoids using the TrapezoidalSum function:

TrapezoidalSum[f(x), 0, 8, 3]

which gives us an approximation of 21.6562. If we keep adding the


amount of trapezoids, then we’ll get closer to the area under the curve. Us-
ing the integral function, we find that the area under the curve is 20.20746
(Table 4.3).
As a numerical technique, the trapezoid rule does a better job of get-
ting us to the area faster. In the context of integrating numerically, certain
methods will yield smaller error—which is a major factor in estimation
that we’d like to minimize. Just to illustrate how quickly the trapezoidal
rule converges to the true area, the scatterplot in Figure 4.23 should drive
the point home.

35
30
Area under the curve

25
20
Lower Riemann Sums
15
Upper Riemann Sums
10 Trapezoidal Rule
5
0
1 2 3 4 5
Step

Figure 4.23.  Comparison of sums of areas under the curve (Steps 1 to 5


correspond to the steps in Table 4.4).
116  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Just like in FEM, your elements don’t need to be squares, rectanges,


or whatever. The shape of the element is entirely dependent on the content.
The elements aren’t required to be 2-dimensional either; they can be sim-
ple lines or curves—even volumes (3D shapes). Figure 4.24 gives a few
examples of different elements.

Figure 4.24.  Examples of elements.

While integration is a simplified example of a numerical technique,


a more engineering style problem deals with the idea of heat transfer.
When the equation for heat transfer of some body is derived, it ends
up being a partial differential equation—which is outside our scope.
Unlike normal differential equations, these involve multiple functions
and their derivatives. Due to this, there are quite a few extra details to
be concerned with. While knowing the fundamentals of engineering are
important to inform basic principles of designing, analysis of structures
in the context of civil engineering or some device in mechanical or elec-
trical engineering could have dozens of equations (not just algebraic
equations either—partial and ordinary differential equations too). That
sounds miserable to solve by hand, and you’re right, it would be. Impos-
sible would be another good adjective to describe the experience, as the
chance of finding some nice solution is slim to none. That’s where FEA
comes in.
Say we had some pole that’s made up of some arbitrary type of steel
and we subject it to a variety of forces, like shear above. Our job is to mea-
sure the flexural behavior of the steel when the forces are applied. How
do we even measure something like that? Well, we could just measure the
displacement on top and bottom and call it day. We can do better than that
though! We discretize the pole and break it into pieces. Figure 4.25 shows
one unit of the discretized beam. The dots at each corner of the cube are
nodes while the faces of the cube are elements. If we wanted to measure
the deflection of the beam in-depth, we can check the displacement at
each node. This would all be done by the computer, leaving it up to you to
interpret the results.
Analyzing Failure in Systems   •   117

Figure 4.25.  Shearing of a cylindrical object.

Looking at the element itself, we can


see the effect the forces had in forcing each
element and the four nodes associated with
it (Figure 4.26). The original element has the
dashed lines and the arrows just show the
direction the nodes shifted. Each node didn’t
necessarily have to shift the same distance Figure 4.26.  Shearing at
the element level.
either, so don’t let this idealized model fool
you! That’s the point of breaking a structure into tiny pieces to analyze, to
find this kind of information out!
Most real-world engineering problems can be solved using any
kind of software that either comes packaged with an oversimplified or
“dumbed-down” FEA package or a stand-alone program that captures the
full power of FEA. While this may seem trivial, FEA is one of the most
powerful methods of solving complex physical engineering problems and
plays a large role in design. In fact, the size of the problem doesn’t matter.
Most huge problems in engineering are tailored nicely for an application of
FEA, allowing us to analyze a model with hundreds of nodes and elements.
While we will not be going into the specifics of FEM, we will be
looking at problems that do use FEA as the primary technique to solv-
ing differential equations and modeling phenomena. If there’s anything
to take away, understand the value and potency that this method brings to
the table—breaking down a giant and complex problem into tiny chunks
to arrive directly at a numerical solution. One easy way to visualize the
solution is a “heat map” which allows us to quickly spot troublesome val-
ues—high stress in this context (Figure 4.27).
At last, we arrive at the FEA Model of the U10 Gusset Plate. The
NTSB model sweeps the estimated stress in the original design, after
118  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Figure 4.27.  Idealized FEA solution.


Source: National Transportation Safety Board report on the Failure of the
I-35 ­Highway Bridge. (www.ntsb.gov/investigations/AccidentReports/Pages/
HAR0803.aspx)

modifications in 1977 and 1998, and the stress at the time that the bridge
­collapsed. Some basic surface analysis of the models begins with pointing
out the meaning of the colors (or shades in black & white). Anything in
white is well below the tolerated stress, shades of gray signify that the
stress is much higher (but still manageable), and black occurs wherever
the internal forces exceed the yield stress.
Looking at the original model, there was already a section of the U10
plate that had internal forces exceeding the yield strength, beginning de-
formation. The modifications from 1977 and 1998 significantly increase
the area of the “danger zone.” Increases in weight from the modifications
did not help the U10 plate’s structural integrity one bit. Fast-forward to
the night of the collapse—the FEA model estimates that the bridge’s U10
plate underwent an increase in internal forces. This increase was large
enough to create specs of black in the graphic, indicating that the forces
inside of the gusset plate were far beyond the allowable amount—causing
the plate to fail.

4.3  DESIGN OF AN ELECTRICAL SYSTEM

Say we have some electric circuit and we would like to know the effective
value of the AC current (which is periodic, aka a function). All that means
is the following: we’re looking for a DC current (a constant value) that will
supply the same average power to a load (which could be a resistor) just
like the AC current. What about voltage, does it have an effective value?
Yes, voltage that changes over time does have an effective value as well. In
fact, a general periodic signal has an effective value.
Here’s the big question: how do we define the effective value of a
periodic function? The answer is quite a mouthful: the effective value is
Analyzing Failure in Systems   •   119

AC VERSUS DC

We see these terms whenever we deal with electricity. DC, or Direct


Current, is a steady flow of current in one direction. Adding two AA
batteries (each 1.5 V) to a flashlight give us 3 V, producing direct cur-
rent through a lightbulb. Power supplies inside computers, car batteries,
and TV remote controls all use DC.
AC, or Alternating Current, is found in any wall outlet. AC means
that current changes directions; in household power, this happens 60
times per second (60 Hz). AC gives us an “effective voltage” or RMS
value as shown. When we plug in electronics, this AC is often converted
to DC: the adapter or transformer you find for many electronic devices
does this conversion.
Why not just use DC everywhere? Transmitting AC power over vast
distances leads to far less power loss than transmitting DC. Further, it
is easier to convert AC power to any DC value than to transform a DC
voltage to a different DC voltage.

This horizontal line is what


we’re looking for, the effective
value of periodic function.

Here is the function,


which is periodic like
sine or cosine.

Figure 4.28.  AC voltage and the RMS, or effective, value.

the square root of the mean of the square of the periodic function. This
definition gives us the other common name for the effective value, the
root-mean-square value (or rms value for short, Figure 4.28).
Now, how does this look in a formula? Let’s start by moving backward
and starting with the square part. That’s pretty easy, we need to square the
periodic function, x (t ).

 x ( t ) 
2
120  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Done. Now, we need to take care of the mean part. Mean is just the
fancy way to say average. For this particular problem we need to find
the average value of a function. Remember how to determine the average
value (or mean value) of a function?

1 b
f avg = ∫ f ( x )  dx
b−a a

We know the f ( x ), but what about the interval? Typically, we’d like to
know the rms value for some period (or interval of time). Period is usually
given as T, so our interval is from the start time, 0 to T.

1
[ x ( t )]   dx
T 2
f avg =
T −0 ∫0

Finally, the root part. To put the cherry on top of this derivation, all
that’s left is to take the square root of what we have.

1
∫0 [ x (t )]
T 2
f avg =   dx
T

Therefore, we define the rms value for a general function, x (t ), over


the period, T, given by

1
∫0 [ x (t )]
T 2
X rms =   dt  
T

If you want the rms value of the current or the voltage, the changes to
the formula are only superficial. For current, X rms and x (t ) become I rms
and i(t ). Similarly, for voltage, X rms and x (t ) become Vrms and v (t ).
The biggest reason for finding the rms value is being able to calculate
the average power that is dissipated (or used) by a resistor. To do this, we
can using the formula for power, but with the rms flavor added in:

2
Vrms
Pavg = I rms
2
R=
R
Analyzing Failure in Systems   •   121

Example 4.4: Finding the rms value of a periodic function and finding
the average power

(1) Find the rms value of the current, We’re given our problem
i ( t ) = sin( t ), over a period of 5 statement.
seconds. Then find the average
power dissipated by a 100 Ω resistor.
(2) 1 T It’s clear that this formula
  [ i ( t )]   dt
2

T ∫0
I rms = will come into play. We
only need to identify the
function and the period.
(3) 1 5 We know that the function
  ∫ [ sin(t ) ]   dt
2
I rms = is sin(t ) and the period
5 0
is 5 seconds.
(4) 1 Looks like a sine squared
sin 2 ( t ) =
2
(1 − cos ( 2t )) popped up. Luckily, we
know a trig identity to
resolve this.
(5) 1 51 Now we can replace
I rms =
5 ∫0 2
(1 − cos ( 2t ))  dt the sine squared with
something we can
evaluate.
(6) 1 5 1 5 Two things happened here,
dt −   ∫ cos ( 2t )  dt
10 ∫0
I rms = we factored out the 1/2
10 0
and made the outside 1/10
as a result, then we split
up the integral into two.
(7) 1 1 5 The first integral is easily
I rms =  ( 5 − 0 ) − ∫ cos ( 2t )  dt done, but the second
10 10 0
integral requires a bit
more effort.
(8) let   u = 2t ,   du = 2   dt We need to use a
substitution, so let u be
the insides of cosine.
1 1 10 Putting the substitution in
cos ( u )  du
2 20 ∫0
I rms = − place, we need to account
for the extra 2 by pulling
out a 1/2. Don’t forget to
change the bounds.
122  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

(9) 1 1 Evaluate the integral


I rms = −
2 20
(cos (10) − cos (0)) and we’ll end up with
this mess. Simplifying
brings this down to a
more manageable value.
(10) I rms = 0.707 A We find that the rms
value for the current is
0.707 A.
(11) P = I 2 R For part 2, we need to
avg rms
calculate the average
power. We now have
Pavg = ( 0.707 ) (100)
2
the rms current and we
know the value of the
resistor, 100 Ω.
(12) Pavg = 49.9849 W After that calculation,
we find that the resistor
yields an average of
49.9849 W given the
current we used.

***

4.4  A REAL-WORLD PROBLEM: A KNEE


REPLACEMENT

Diseased knee joints that cause substantial amounts of pain are excel-
lent candidates for knee replacements. We’ve come a long way since the
first knee arthroplasty, which was the equivalent of nailing a hinge into
your bones and hoping for the best. Today, the surgeon slices away at any
cartilage and bone deemed to be damaged or diseased and replaces the
removed bones with an artificial joint. The joints used as a replacement
for your knee are made of high-grade materials. For anything that should
be stiff and hard, metal alloys do the job. If something needs to be bendy
and stretchy, plastics and polymers are used.
You may have noticed during your life experiences that your body
parts don’t necessarily move in every direction possible, unless you’re
a contortionist. This is mainly due to the structure of the knee itself.
Whenever you do something that involves your knee moving, you can
thank cartilage. This elastic tissue allows the surfaces of the joint to
get cozy and slide across each other without difficulty—that is hyaline
cartilage. Whenever your knee is under pressure, the fibrous cartilage
Analyzing Failure in Systems   •   123

Figure 4.29.  Knee flexion angle.

provides the necessary resistance. The menisci also play a role in protect-
ing bones from rubbing together and act as shock absorbers. Imagine all
of the stress your knee goes under over the course of the day, especially
if you’re athletic. With bones constantly rubbing up against each other, it
makes sense that the tissue will deteriorate over time. If that wasn’t bad
enough, this cartilage isn’t a fan of regeneration. Any new tissue that
does form will be of far lesser quality than what you started with. Your
limited knee movement can be blamed on the ligaments. Think of these
like exceptionally strong rubber bands that can only stretch a limited
amount, and help bring the knee back to its original position. These keep
your knee joints stable by mostly restricting motion. One main concern
of complete knee replacement is the loss of flexibility. More traditional
knee replacements have trouble allowing the patient to flex beyond
an angle of 115° (Figure 4.29). This is more crucial for high-demand
patients like athletes.
Table 4.4 lists 15 patients whose knees were tested to determine the
maximum range of motion—the largest angle possible the knee could bend.
After performing some statistical analysis on that sample of pre- and
postoperation cases, you may or may not have noticed that patients expe-
rienced a wide variety of results. Sure, the differences can be caused by a
number of conditions, but the unfortunate reality is the fact that precision
isn’t a guarantee. Unlike some materials like steel, the various properties
we associate with metals don’t necessarily hold true for biological tissue.
For example, if we set up a bar steel in a tension test (where we slowly pull
the bar apart by gripping it at each end until the bars snaps in half), run the
test, rinse and repeat with more steel bars, then we’d get consistent results.
In fact, we have all sorts of diagrams and databases that tell us all the
material properties that we’d ever want to know. On the other hand, if we
somehow were in (legal) possession of a bunch of ligaments from human
knees, we can’t exactly deduce that we’d get consistent results. Sure, a
124  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Table 4.4.  Comparison of patient’s flexion angle pre- and


postreplacement

Patient Preoperation (angle) Postoperation (angle)


1 68 94
2 80 115
3 79 112
4 63 126
5 98 134
6 125 138
7 110 140
8 93 145
9 88 104
10 115 134
11 92 127
12 128 136
13 112 140
14 105 128
15 120 137

ligament is a ligament, but who


did it come from? How healthy
were they? How old? Surely we
can imagine what we’d need to
know the life story of each sam-
ple; therefore, we tend to settle
for a range of values.
Figure 4.30.  Representation of the divot
The same ideas apply for the
caused by degradation.
cartilage that recedes or degrades
in the knee from different diseases. Say we had the knee of an athlete whose
sport has brewed a perfect storm that allowed cartilage to erode and form a
divot (circled). Due to the severity of the degradation, our patient has been
experiencing severe discomfort and is slowly losing the ability to compete.
The abstraction of that situation is shown in Figure 4.30.
Using some engineering design thinking, we could feasibly put pen to
paper and create a model to fill that divot. Our issue is how to calculate the
Analyzing Failure in Systems   •   125

Max thickness = 3 mm

Figure 4.31.  Identifying the maximum thickness.

maximum depth that our replacement cartilage needs to go and the area of
the cross-section (the striped piece in Figure 4.31).
Note that when we add metal to our body during an operation similar
to a knee replacement, there’s always a possibility that the body will reject
the new material. Under ideal conditions, we could use whatever we want
to fix the problems in our bodies, but that’s not the case.
Stainless steel was the king of implants for dentistry and general sur-
gery until the 1980s when titanium stole the crown for its high strength to
weight ratio. Yet, a few difficulties in manufacturing and expenses popped
up, limiting its application to a degree. Don’t try aluminum though! It can-
not possibly come to the level of stainless steel or titanium even if it tried.
We tend to keep any aluminum parts outside of the body since the chance
that it won’t play well with the human body is a bit questionable.
What about plastics? They are malleable (can be pressed out of shape
without breaking), are able to change shape, and are cheap to produce. To
provide some context, Table 4.5 contains common materials often used in
knee replacements.

Table 4.5.  Summary of materials common in knee replacements

Material Advantages Disadvantages


Stainless steel Suited for temporary replace- Limited ability to
ments, proven biocompatible withstand corrosion
Cobalt- All around tough metal, corro- Small possibility of
chromium alloy sion resistant, biocompatible allergic reaction to
special metals
Titanium Used for situations where high
strength is not necessary, inert,
low density, more natural
Tantalum Extremely flexible, corrosion
resistant, aids in bone remodeling
Zirconium alloy Believed to be long lasting, can
be lubricated
126  •   THE APPLICATION OF MATHEMATICS IN THE ENGINEERING

Now, consider how we would design a knee replacement for the ath-
lete whose knee has atrophied. How would we determine which materials
to use for the athlete’s needs? What range of motion does the athlete need?
What are the tolerances on the replacement’s dimensions?
Take some time to sit down and think about how a knee replacement
would be designed given this short primer on what is involved in the process.

4.5  CLOSING THOUGHTS

Wait, not a single equation has appeared in the last section. Numbers sure,
but no complex differential equations or formulas. We also did not give
any examples and instead made some asks. What’s the big deal? Contrary
to what the engineering courses laden with equations would lead one to
believe, engineering practice is more concerned with design as opposed
to what we’ve engaged in up to this point—engineering analysis. Sure,
modeling has its place, but mathematical principles will guide our design
processes in an ancillary manner. In practice, we will be researching, pro-
totyping, testing, and refining until we reach a suitable solution to the given
problem. Mathematics is often wrongly presented as a cold, uncreative tool
to drive the work of multiple professions in the workforce. One fallacy is
the existence of a single correct answer for every problem, which is simply
not true. The same is true for problems in the “wild,” on the job. Rather
than attempting to find the only possible solution to solve the problem,
we are attempting to find a feasible solution that satisfies the constraints
we are given—this is the essence of ill-structured problems (Figure 4.32).
Sadly, life’s problems are not in the back of the book, and in the prac-
tice of engineering, we all can construct something from inside the set of
feasible solutions—someday.

Solution Space

Set of Feasible Solutions


Meeting all Constraints

Infeasible Solutions

Figure 4.32.  An illustration of an “ill-structured”


problem.
About the Authors

David Reeping is a graduate student pursuing a PhD in engineering


­education at Virginia Tech and is an NSF graduate research fellow. He
received his BS in engineering education with a mathematics minor from
Ohio Northern University. He was a Choose Ohio First scholar inducted
during the 2012–2013 school year as a promising teacher candidate in
STEM. He was the recipient of the Remsburg Creativity Award for 2013
and the ­DeBow Freed Award for outstanding leadership as an undergradu-
ate student (sophomore) in 2014. He is also a member of the mathematics,
education, and engineering honor societies: Kappa Mu Epsilon, Kappa
Delta Pi, and Tau Beta Pi, respectively.
He has extensive experience in curriculum development in K-12 and
creates material for the Technology Student Association’s annual TEAMS
competition. David has coauthored two books related to engineering—
Principles of Applied Engineering for Pearson-Prentice Hall and Introduc-
tory Engineering Mathematics for Momentum Press. His research interests
include model/method transferability, threshold concepts to inform curric-
ulum development, information asymmetry in higher education processes
(e.g., course articulation), and issues in first-year engineering.

Kenneth Reid is an associate professor and assistant department head for


Undergraduate Programs in Engineering Education at Virginia Tech. He
earned his PhD in engineering education in 2009, the seventh in the nation.
He and his coauthors were awarded the William Elgin Wickenden Award
for 2014, recognizing the best paper in the Journal of Engineering Educa-
tion and the best paper award for the Educational Research and Methods
(ERM) Division of ASEE in 2014. He was awarded an IEEE-USA Pro-
fessional Achievement Award in 2013 for designing the nation’s first BS
degree in engineering education. He is active in engineering within K-12,
serving on the TSA Board of Directors. His research interests include
128  •   ABOUT THE AUTHORS

s­ uccess in first-year engineering, engineering in K-12, introducing entre-


preneurship into engineering, and international service and engineering.
He has written books on introductory engineering and digital electronics,
has taken multiple teams of students to the Dominican Republic through
Solid Rock International, and has a number of teaching awards.
Index

A Determinants, 19–23
Airport locations, mathematical Differential equations
model of, 3–4 convolution, 51–53
Algebraic equation, 34–35 defined, 29
Alternating current (AC) vs. direct integral transforms, 40–41
current (DC), 119 Laplace transform, 41–51
Aluminum and alloys, 101 solutions to, 31
American Association of State “typical”, 29–39
Highway Officials (AASHO), Differential operator, 30–31
102, 103 solving higher order equation
American Society of Civil using inspection, roots of,
Engineers (ASCE), 90 34–35
Angular frequency, 83 Direct current (DC), alternating
Axial force, finding, 96–98 current (AC) vs., 119
Distributed load, 105
B nonlinear, 112–113
Bridge 9340. See I-35 West replacing, 109–113
Mississippi Bridge Dot product, 16
Ductile iron, 101
C
Cables, modeling of, 83–89 E
Cascade connection, 77–79 Elastic modulus. See Young
“Centered at a”, meaning of, 5 modulus
Centroid, 106–109 Elasticity, 98
Complementary solution, 39 Electrical system
Concentration, 108 design of, 118–122
Control engineering, 80 Laplace transforms, 48–51
Controls, concept of, 76–80 Elementary functions, 9
Convolution, 51–53, 58 Elementary row operations,
Cramer’s rule, 23–28 13, 14
Engineering mathematics. See
D Systems, and mathematics
Delta function, 42 Equilibrium, requirements for,
Derivative operator, 30–31 90–98
130  •   Index

F I
Failure analysis I-35 West Mississippi Bridge,
electrical system, design of, 89–90
118–122 structural loads, 105–113
I-35 West Mississippi Bridge, U10 plate
89–90 finite element analysis of,
finite element analysis of U10 113–118
plate, 113–118 material properties of,
material properties of U10 98–101
plate, 98–101 requirements for equilibrium
requirements for equilibrium in in, 90–98
U10 plate, 90–98 strength of materials, 101–105
strength of materials, U10 Identity matrix, 18–19
plate, 101–105 Impulse function, 42
structural loads, 105–113 Impulse response, 55–59
real-world problem, knee Initial value problems, 32–33
­replacement, 122–125 Integral transforms, 40–41
Tacoma Narrows Bridge, 81–82
cables, modeling of, 83–89 K
simple harmonic motion, 83 Knee replacement, 122–125
Failure by shear, 100 materials common in, 126
Fibrous cartilage, 122–123
Finite element analysis (FEA), L
113–118 Laplace, Pierre-Simon, 41–42
Forced factoring, 46 Laplace transforms, 41–44
Fracture, 100 electrical system, 48–51
Free body diagram, 82 mechanical system, 47–48
Free coefficients, 25–27 more difficult, 45–47
Frequency, 83 simple, 44–45
Fundamental Theorem of Ligaments, 123–124
Calculus, 112 Linearity, 44, 46
G
Galloping Gertie. See Tacoma M
Narrows Bridge Maclaurin series, 5–10
Gauss–Jordan Elimination Marginal stability, 74–76
method, 14 Mass spring damper system,
General solution, 31–32 65–74
GeoGebra, 114 MATLAB, 12, 82
Graph. See Network Matrices, defined, 10
Gusset plate. See U10 plate Matrix multiplication, 15–18
properties of, 18–19
H Mechanical system, Laplace
Harmonic motion, 83 transforms, 47–48
Homogeneous differential Menisci, 123
equation, 31 Modeling systems
Hyaline cartilage, 122 of airport locations, 3–4
Index  •   131

Cramer’s rule, 23–28 Resultant shear, 102–103


determinants, 19–23 Root-mean-square value, 119–122
matrices to solve problems, Roots, of differential operator, 34–37
10–19 Row echelon form, matrix, 14
model Row expansion method, 20
defined, 1 Rube Goldberg machine, 77
reasons for making, 1–2
note on creating, 2–3 S
Taylor and Maclaurin series, Series connection. See Cascade
5–10 connection
Multiple systems, 76–80 Shear, failure by, 100
Shear strength, 101
N Shear stress, 101
National Transportation Safety calculation of, 104
Board (NTSB), 90 Simple harmonic motion, 83
Network, 3, 4 Simple shear, 101
Nodal analysis, 27 Stability, 60–76
Nonhomogeneous differential Stabilization, 79
equation, 30, 37–39 Steel, 101
Numbers, meaning of, 12 Stress, 89
Stress–strain curve, 100
O Structural loads, 105–113
Overlapping poles, 74–76 Systems, and mathematics
connecting multiple systems
P and concept of controls,
Parallel connection, 77–78 76–80
Partial fractions, 68 failure analysis in. See Failure
Particular solution, 31–32 analysis
Pendulum, 33 finding output of, 59
Poisson’s Ratio, 99 pole-zero plot for. See Pole-zero
Pole-zero plot, 61, 63–65, 68, 70 plot
defined, 61 stability, 60–76
for hypothetical system, 64 transfer function and impulse
for marginally stable system, 65 response, 55–59
for mass spring damper, 68,
70, 72 T
for system, 63 Tacoma Narrows Bridge, 81–82
Principle of superposition, 55 cables, modeling of, 83–89
simple harmonic motion, 83
R Taylor series, 5–10
Radius of convergence, 6, 8 Transfer function, 55–59
Ratio test, 6 poles and zeroes of, 60–61
Rectangular load, 109 Transform table, 45
Reduced row echelon form, Trapezoidal rule, 114
matrix, 14 TrapezoidalSum function, 115
132  •   Index

Triangular load, 109 V


Typical differential equations, Voltages in circuit, matrices and, 27
29–39
W
U Washington Toll Bridge
U10 plate Authority, 82
finite element analysis of,
113–118
free body diagram of, 92 Y
material properties of, 98–101 Yield strength, 100–101
requirements for equilibrium in, Yield stress, 89
90–98 Young modulus, 98–99
strength of materials, 101–105
Ultimate strength, 100–101 Z
Unit step function, 42 Zero input solution, 55
Unsimplifying process, 46 Zero state solution, 55
OTHER TITLES IN OUR GENERAL ENGINEERING AND
K-12 ENGINEERING EDUCATION COLLECTION
John K. Estell, Ohio Northern University and Kenneth J. Reid, Virginia Tech, Editors

• Lean Engineering Education: Driving Content and Competency Mastery by Shannon


Flumerfelt, Franz-Josef Kahlen, Anabela Alves, and Anna Bella Siriban-Manalang
• Sustainable Engineering by Kaufui Vincent Wong
• Introductory Engineering Mathematics by David Reeping and Kenneth Reid
• Cracking the Code: How to Get Women and Minorities into STEM Disciplines and
Why We Must by Lisa M. MacLean
• Engineering Design and the Product Life Cycle: Relating Customer Needs, Societal
Values, Business Acumen, and Technical Fundamentals by Kenneth J. Reid and
John K. Estell

Momentum Press offers over 30 collections including Aerospace, Biomedical, Civil,


Environmental, Nanomaterials, Geotechnical, and many others. We are a leading book
publisher in the field of engineering, mathematics, health, and applied sciences.

Momentum Press is actively seeking collection editors as well as authors. For more
information about becoming an MP author or collection editor, please visit
http://www.momentumpress.net/contact

Announcing Digital Content Crafted by Librarians


Concise e-books business students need for classroom and research

Momentum Press offers digital content as authoritative treatments of advanced engineering


topics by leaders in their field. Hosted on ebrary, MP provides practitioners, researchers,
faculty, and students in engineering, science, and industry with innovative electronic content
in sensors and controls engineering, advanced energy engineering, manufacturing, and
materials science.

Momentum Press offers library-friendly terms:


• perpetual access for a one-time fee
• no subscriptions or access fees required
• unlimited concurrent usage permitted
• downloadable PDFs provided
• free MARC records included
• free trials

The Momentum Press digital library is very affordable, with no obligation to buy in future
years.

For more information, please visit www.momentumpress.net/library or to set up a trial in the


US, please contact mpsales@globalepress.com.
EBOOKS The Application of Mathematics

REEPING • REID
FOR THE in the Engineering Disciplines GENERAL ENGINEERING AND K-12
ENGINEERING David Reeping • Kenneth J. Reid ENGINEERING EDUCATION COLLECTION
LIBRARY This text serves as the companion text to Introductory John K. Estell and Kenneth J. Reid, Editors
Create your own Engineering Mathematics, which introduces common
mathematical concepts we see in engineering, including
Customized Content
trigonometry, calculus, and functions. This text assumes a
Bundle—the more level of mathematics of a high school senior, plus some ele-
books you buy, ments from the introductory text. Additional concepts we
the greater your
discount!
see in engineering are also introduced: specifically, matri-
ces, differential equations, and some introduction to series.
After, the basics of system analysis, including stability, are
The Application
of Mathematics

The Application of Mathematics in the Engineering Disciplines


explored.
THE CONTENT The concepts are presented using examples rather than
• Manufacturing strict mathematical derivation. As a result, this text likely will
not be an effective substitute for a differential equations

in the Engineering
Engineering
• Mechanical course, but by illustrating the implementation of differential
equations, it can be a companion to such a course. We pri-
& Chemical
marily use historical events as examples (including failures)

Disciplines
Engineering to illustrate the use of mathematics in engineering and the
• Materials Science intersections of the disciplines. We hope you develop an
& Engineering appreciation for how to apply these concepts, and find a
• Civil & new lens through which to view engineering successes (and
Environmental failures).
Engineering David Reeping is a PhD candidate in engineering edu-
• Advanced Energy cation at Virginia Tech and a National Science Founda-
Technologies tion graduate research fellow. David was the recipient of
the Remsburg Creativity Award for 2013 and the DeBow
Freed Award for outstanding leadership as an under-
THE TERMS
graduate student (sophomore) in 2014. He has extensive
• Perpetual access for experience in curriculum development in K-12 and cre-
a one time fee ates material for the Technology Student Association’s
• No subscriptions or annual TEAMS competition.
access fees Kenneth J. Reid is an associate professor in engineering
• Unlimited education at Virginia Tech. He earned his PhD in engi-
concurrent usage neering education in 2009. Among other awards, he and
• Downloadable PDFs his coauthors received the Wickenden award (2014), best
• Free MARC records paper award for the Educational Research and Meth-
ods Division of ASEE (2014) and IEEE-USA Professional
Achievement Award (2013) for developing the nation’s
For further information,
a free trial, or to order,
first BS degree in engineering education. He is active in David Reeping
engineering within K-12, including the Technology Stu-
contact: 
sales@momentumpress.net
dent Association (TSA) board of directors. Kenneth J. Reid

Potrebbero piacerti anche