You are on page 1of 49

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.
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.

You might also like