Engineering Mechanics Statics 1630433197 PDF
Engineering Mechanics Statics 1630433197 PDF
Engineering Mechanics Statics 1630433197 PDF
Engineering Mechanics:
Statics
This is the first of two courses to describe how objects move and
the forces that cause motion. This combines math and physics
fundamentals with real-world application. A structured problem-
solving process is included, and by the end of the book, you should
be able to recognize and describe motion all around you in your
everyday life.
Chapter 1 contains the fundamental math and physics concepts
including vectors, Pythagorean theorem, sine and cosine laws, dot
product, Newton’s laws, weight and mass, unit conversions, and the
problem solving process.
Chapter 2 explains the difference between particles and rigid
bodies and introduces free-body diagrams and equilibrium
equations for particles.
Chapter 3 contains introductory rigid body concepts, including
cross products, the right hand rule, torques/moments and couples,
distributed loads and reaction/support forces.
Chapter 4 introduces free-body diagrams and equilibrium
equations for rigid bodies, as well as external forces, frictional and
impending motion.
Chapter 5 introduces trusses and two methods to solve truss
systems: method of joints and method of sections.
Introduction | 1
Chapter 6 explains internal forces and breaks down shear/
moment diagrams.
Chapter 7 introduces center of mass, mass moment of inertia,
area moment of inertia, and the parallel axis theorem.
Appendix A has a reference list of open textbooks.
Key Takeaways
2 | Statics
But that’s not enough. Look at the solutions that are posted and
practice. You’ll get out of it what you put in, and statics can be fun.
It’s how engineers apply physics concepts to the real world. I hope
you learn to love learning engineering as much as I enjoy teaching
it. Hopefully, you’ll get a sense of the wonder of engineering through
this book.
Introduction | 3
CHAPTER 1:
FUNDAMENTAL CONCEPTS
Static vs Dynamic Motion
Before we start, what is the difference between static and
dynamic? Static problems are all problems where there is no
acceleration. As you’re driving down the road and you’re cruising
along at a constant velocity, that is a static problem. As soon as you
start to slow down for a stop light or speed up, you are in dynamic
motion, and that’s much more complicated. For this course, we
will only consider problems where there is no motion (such as the
‘static’ is used in the English language), or constant velocity. Be
prepared to give many examples of static versus dynamic, because
before you can solve a problem, you have to know what type of
problem it is!
Introduction to Chapter 1: Fundamental Concepts
This chapter has a lot of concepts from math and physics that
are necessary for you to understand before we can apply them in
engineering contexts. It’s kind of like: before you can write an essay
to express your opinion, you need to know how to write the a,
b, c’s and what each word means. Here, you need to know how
to compute a cross product before you can calculate how much
Torque is created from a force.
Some of this might be new. Some of this might be familiar, but we
might be applying it in a different way. (Such as calculating torque –
this is not what you learned in high school physics!) Some of it might
feel new, so practice practice practice!
Here are the sections in this Chapter:
Here are the key equations and concepts you will learn in this
chapter
6 | Statics
1.1 Preparatory Concepts
8 | Statics
Source: University Physics Volume 1, OpenStax CNX
https://courses.lumenlearning.com/suny-
osuniversityphysics/chapter/2-1-scalars-and-vectors
Key Takeaways
1st Law:
10 | Statics
Source: Engineering Mechanics, Jacob Moore et al.,
http://mechanicsmap.psu.edu/websites/
1_mechanics_basics/newtons_first_law/firstlaw.html
2nd Law:
3rd Law:
12 | Statics
Since all forces are exerted by bodies (either directly or
indirectly), all forces come in pairs, one acting on each
of the bodies interacting.
Application: 1st law: a rock rolling down the hill will keep
going unless it hits a tree. 2nd law: the amount of forces on
the rock and how massive (heavy) it is will determine how
much it is accelerating (or decelerating). 3rd law: the rock is
pushing on the ground with the same amount of force as
the ground is pushing on the rock, but in the opposite
direction.
14 | Statics
1.1.3 Units
16 | Statics
define the SI base units. The following table lists these
seven ISQ base quantities and the corresponding SI base
units.
SI
ISQ Base
Base
Quantity
Unit
Meter
Length
(m)
Kilogr
Mass
am (kg)
Secon
Time
d (s)
Electrical Ampe
current re (A)
Thermodyna Kelvin
mic temp. (K)
Amount of Mole
substance (mol)
Luminous Cande
intensity la (cd)
18 | Statics
https://courses.lumenlearning.com/suny-
osuniversityphysics/chapter/1-2-units-and-standards/
m (meter), km
ft (foot), mi (mile), in
Length (kilometer), mm
(inch)
(milimeter)
Key Takeaways
20 | Statics
The first thing to do is to list the units you have and the
units to which you want to convert. In this case, we have
units in meters and we want to convert to kilometers.
Next, we need to determine a conversion factor relating
meters to kilometers. A conversion factor is a ratio that
expresses how many of one unit are equal to another
unit. For example, there are 12 in. in 1 ft, 1609 m in 1 mi,
100 cm in 1 m, 60 s in 1 min, and so on. In this case, we
know that there are 1000 m in 1 km. Now we can set up
our unit conversion. We write the units we have and
then multiply them by the conversion factor so the units
cancel out, as shown:
[latex]80m=8.0\times10^1m=8.0\
times10^{-2}km=0.080km [/latex]
[latex]80m=8.0\times10^1m=8.0\
times10^{-2}km=0.080km [/latex]
Going back and forth between SI and English will become very
useful skill. If you can memorize km to mi, ft to m, and inches to
ft, you’ll be able to communicate better with coworkers. Here are
common conversions you’ll need for this course:
Force N lb 1 lb = 4.448 N
Pressure Pa psi
1psi = 6895 Pa
All of the other units that we will encounter will be a mix of these
units (intensity w = N/m or lb/ft). One additional conversion that is
common is 1 lb = 2.2 kg, though this only works on Earth because it
is mixing kg and lb (see next section). For a full table, MechanicsMap
has a pdf available: http://mechanicsmap.psu.edu/websites/
UnitConversion.pdf
22 | Statics
Units outside the SI that
are accepted for use with
the SI
Symb
Name Value in SI units
ol
1 eV = 1.602 18 x 10-19 J,
electronvolt (d) eV
approximately
Source: https://www.physics.nist.gov/cuu/Units/
outside.html
24 | Statics
Key Takeaways
[latex]g(r)=G\frac{M}{r^2}[/latex]
26 | Statics
where G = 6.67 × 10−11 is Newton’s constant of gravity,
and r is the distance from the centre of the object.
In the English language, the words ‘mass’ and ‘weight’ are used
interchangeably. A person might say, “I weigh 50 kg”, but in statics
language, that’s wrong! Or more accurately, that language isn’t
precise enough for statics.
28 | Statics
While most Canadian companies use SI units, it’s important to be
familiar with English, so you should learn slugs. You don’t want to be
excluded from a conversation at your future job.
Note lbm (pound-mass) is not used in this book, though some
textbooks use it as a mass value. When lb is used, it is assumed to be
lbf (pound-force).
Key Takeaways
30 | Statics
Key Takeaways
32 | Statics
0px;padding: 0px;margin: 0px">in terms of the
coordinates x xx" role="presentation" style="overflow:
initial;font-style: normal;font-weight: normal;line-
height: normal;font-size: 14px;text-indent: 0px;text-
align: left;text-transform: none;letter-spacing:
normal;float: none;direction: ltr;max-width: none;max-
height: none;min-width: 0px;min-height: 0px;border:
0px;padding: 0px;margin: 0px">and y.y.y."
role="presentation" style="overflow: initial;font-style:
normal;font-weight: normal;line-height: normal;font-
size: 14px;text-indent: 0px;text-align: left;text-
transform: none;letter-spacing: normal;float:
none;direction: ltr;max-width: none;max-height:
none;min-width: 0px;min-height: 0px;border:
0px;padding: 0px;margin: 0px">
$$\sin\theta=y\;\;\;\csc\theta=\frac{1}{y}\\\cos\
theta=x\;\;\;\sec\theta=\frac{1}{x}\\\tan\
theta=\frac{y}{x}\;\;\;\cot\theta=\frac{x}{y}$$
If x=0x=0,secθx=0,secθ" role="presentation"
style="overflow: initial;font-style: normal;font-weight:
normal;line-height: normal;font-size: 14px;text-indent:
0px;text-align: left;text-transform: none;letter-spacing:
normal;float: none;direction: ltr;max-width: none;max-
height: none;min-width: 0px;min-height: 0px;border:
0px;padding: 0px;margin: 0px">, secθ and tanθ are
undefined. If y=0, then cotθ and cscθ are undefined.
34 | Statics
0px;text-align: left;text-transform: none;letter-spacing:
normal;float: none;direction: ltr;max-width: none;max-
height: none;min-width: 0px;min-height: 0px;border:
0px;padding: 0px;margin: 0px">θ, the coordinates x and
y satisfy:
[latex]x=r\cos\theta[/latex]
[latex]x=r\sin\theta[/latex]
Trigonometric Identities
36 | Statics
Source: Calculus Volume 1, Gilbert Strang & Edwin
“Jed” Herman, https://openstax.org/books/calculus-
volume-1/pages/1-3-trigonometric-functions
Key Takeaways
38 | Statics
1.2 XYZ Coordinate Frame
We need a standard to be able to share a common language. The
Cartesian coordinate frame lets us express the location of a point so
that others can understand what we’re talking about.In this section,
we’ll look at 2d and 3d coordinate frames.
40 | Statics
It is customary to denote the positive direction on
the x-axis by the unit vector i and the positive direction
on the y-axis by the unit vector j. Unit vectors of the
axes, i and j, define two orthogonal directions in the
plane. As shown in the figure above, the x– and y–
components of a vector can now be written in terms of
the unit vectors of the axes:
42 | Statics
of a vector (where b stands for “beginning”) and the
coordinates e(xe, ye) of the end point of a vector
(where e stands for “end”), we can obtain the scalar
components of a vector simply by subtracting the origin
point coordinates from the end point coordinates:
44 | Statics
normal; font-weight: normal; letter-spacing: normal;
float: none; direction: ltr; max-width: none; max-height:
[latex]A_z=z_e-z_b[/latex]
[latex]A=\sqrt
{A_{x}^{2}+A_{y}^{2}+A_{z}^{2}}[/latex]
46 | Statics
Source: University Physics Volume 1, OpenStax CNX,
https://courses.lumenlearning.com/suny-
osuniversityphysics/chapter/2-2-coordinate-systems-
and-components-of-a-vector/
Key Takeaways
48 | Statics
1.3 Vectors
1.3 Vectors | 49
\sqrt{{\underline{\hat{i}}}^{2}+{\underline{\hat{ j}}}^{2}+{\un
derline{\hat{k}}}^{2}} = 1[/latex]
• The unit vector can be calculated from the magnitude and
vector: [latex]\underline{\hat{a}} =\vec A/|A|[/latex]
In 2d & 3d:
50 | Statics
Source: University Physics Volume 1, OpenStax CNX,
https://courses.lumenlearning.com/suny-
osuniversityphysics/chapter/2-2-coordinate-systems-
and-components-of-a-vector/
1.3 Vectors | 51
1.3.2 Componentizing a Vector
In 2d:
To find the components of a
vector (A) in 2 dimensions (the x
and y portions Ax and Ay), use
SOH CAH TOA:
[latex]\vec A=A_x\
underline{\hat{i}}+A_y\
underline{\hat{ j}}[/latex]
Ax = |A| cos(Θ)
Source: Introductory Physics, Ryan
Ay = |A| sin(Θ) Martin et al.,
|A|2 = Ax2 + Ay2 (magnitude) https://openlibrary.ecampusontario.ca
/catalogue/
tan(Θ) = Ax / Ay (direction) item/?id=4c3c2c75-0029-4c9e-967f-41
f178bebbbb p814
In 3d:
[latex]\vec A=A_x\underline{\hat{i}}+A_y\
underline{\hat{ j}}+A_z\underline{\hat{k}}[/latex]
|A|2 = Ax2 + Ay2+ Az2 (magnitude)
[latex]\begin{aligned} &\hat{a}=\frac{\vec A}{|\vec A|}
\end{aligned}=\frac{{A}_{x} \underline{\hat{\imath}}+A_{y}
\underline{\hat{\jmath}}+{A}_{z}
\underline{\hat{k}}}{\sqrt{\left({A}_{x}\right)^{2}+\left({A}_{y}\ri
ght)^{2}+\left({A}_{z}\right)^{2}}}[/latex]
52 | Statics
A is the position vector. You can add individual position vectors to
find the total position traveled (c = a + b), for example if someone
walks from one point on campus to another, they would rarely walk
in one straight line like c. In the image below, imagine that there
is a building in the square near where a and b meet, so the person
couldn’t take c but had to walk around. The total distance traveled
is |a| + |b|, not |c| (because |c| ≠ |a| + |b|).
Subtraction works the same way, but instead of going from tail to
head of the arrow, the reverse direction is taken, from head to tail.
For example, a = c – b, follow c from tail to head, then go in the
reverse direction of b from head to tail, and you end up at a.
1.3 Vectors | 53
Vectors can be added together and multiplied by
scalars. Vector addition is associative and commutative,
and vector multiplication by a sum of scalars is
distributive. Also, scalar multiplication by a sum of
vectors is distributive:
54 | Statics
This vector equation means we must have
simultaneously [latex]A_x-B_x=0[/latex], [latex]A_y-
B_y=0[/latex], and [latex]A_z-B_z=0[/latex]. Hence,
we can write [latex]\vec A=\vec B[/latex] if and only if
the corresponding components of vectors [latex]\vec
A[/latex] and [latex]\vec B[/latex] are equal:
1.3 Vectors | 55
[latex]\begin{matrix}R_x = A_x+B_x\\R_y =
A_y+B_y\\R_z = A_z+B_z\end{matrix}[/latex]
Key Takeaways
56 | Statics
componentized into x and y, using the angle and the weight
of the person to calculate it.
1.3 Vectors | 57
1.4 Dot Product
A dot product produces a single number to describe the product
of two vectors. If you haven’t taken linear algebra yet, this may
be a new concept. This is a form of multiplication that is used to
calculate work, unit vectors, and to find the angle between two
vectors.
[latex]\vec A\cdot \vec B=|\vec A||\vec B|\cos\theta[/latex]
Dot Product
$$\vec A=A_x\underline{\hat{i}}+A_y\
underline{\hat{ j}}+A_z\underline{\hat{k}}\text{ and
}\vec B=B_x\underline{\hat{i}}+B_y\
underline{\hat{ j}}+B_z\underline{\hat{k}}$$
[latex]\vec A\cdot\vec
B=A_xB_x+A_yB_y+A_zB_z[/latex]
60 | Statics
[latex]\vec A \text{ and }\vec B[/latex] is obtained by
taking the inverse cosine of the expression above.
1. Source: https://en.wikipedia.org/wiki/
Dot_product#/media/File:Dot_Product.svg
Key Takeaways
62 | Statics
Application: Two ropes attached to a sign are being pulled
in different directions. To find the angle between them, use
the dot product of the two vectors.
a×b
b
||a|| ||b||
sin ϴ
n ϴ
a
b×a
The black
arrow is perpendicular to the grey plane made from blue and red
vectors). This is how you will find the amount of torque created from
a force, which we will do many times. Also, unlike the dot product, a
x b is a different direction than b x a.
$$ \vec A\times\vec B=\begin{bmatrix}
\underline{\hat{i}} & \underline{\hat{ j}} & \underline{\hat{k}} \\
A_x & A_y & A_z \\
B_x & B_y & B_z
\end{bmatrix} $$
1. https://commons.wikimedia.org/wiki/File:Cross-
product-with-area.svg
$$\underline{\hat{i}}\times\
underline{\hat{i}}=0\\\underline{\hat{ j}}\times\
underline{\hat{ j}}=0\\\underline{\hat{k}}\times\
underline{\hat{k}}=0$$
66 | Statics
cross product of any one of them with any other one of
them is the product of the two magnitudes, that is, 1.
Now how about the direction? Let’s use the right hand
rule to get the direction of i×j:
• [latex]\underline{\hat{i}}\times\
underline{\hat{ j}}=\underline{\hat{k}}[/latex]
• [latex]\underline{\hat{ j}}\times\
underline{\hat{k}}=\underline{\hat{i}}[/latex]
• [latex]\underline{\hat{k}}\times\
underline{\hat{i}}=\underline{\hat{ j}}[/latex]
• [latex]\underline{\hat{ j}}\times\
underline{\hat{i}}=-\underline{\hat{k}}[/latex]
• [latex]\underline{\hat{k}}\times\
underline{\hat{ j}}=-\underline{\hat{i}}[/latex]
• [latex]\underline{\hat{i}}\times\
underline{\hat{k}}=-\underline{\hat{ j}}[/latex]
[latex]\underline{\hat{\textbf{i}}},\underline{\hat{\t
extbf{ j}}},\underline{\hat{\textbf{k}}},\underline{\hat{
\textbf{i}}},\underline{\hat{\textbf{ j}}},\underline{\hat
{\textbf{k}}}[/latex]
68 | Statics
vector [latex]\vec A[/latex] can be expressed in terms
of unit vectors:
$$\vec{\textbf{A}}=A_x\hat{\underline{i}}+A_y\
hat{\underline{ j}}+A_z\hat{\underline{k}}$$
$$\vec{\textbf{A}}\times\vec{\textbf{B}}=(A_x\
hat{\underline{i}}+A_y\hat{\underline{ j}}+A_z\
hat{\underline{k}})\times(B_x\hat{\underline{i}}+B_y\
hat{\underline{ j}}+B_z\hat{\underline{k}})$$
$$\begin{aligned}\vec{\textbf{A}}\times\
vec{\textbf{B}}=&A_x\hat{\underline{i}}\times(B_x\
hat{\underline{i}}+B_y\hat{\underline{ j}}+B_z\
hat{\underline{k}})+\\&A_y\
hat{\underline{ j}}\times(B_x\hat{\underline{i}}+B_y\
hat{\underline{ j}}+B_z\hat{\underline{k}})+\\&A_z\
hat{\underline{k}}\times(B_x\hat{\underline{i}}+B_y\
hat{\underline{ j}}+B_z\
hat{\underline{k}})\end{aligned}$$
$$\begin{aligned}\vec{\textbf{A}}\times\
vec{\textbf{B}}=&A_x\hat{\underline{i}}\times B_x\
hat{\underline{i}}+A_x\hat{\underline{i}}\times B_y\
hat{\underline{ j}}+A_x\hat{\underline{i}}\times B_z\
hat{\underline{k}}+\\&A_y\hat{\underline{ j}}\times
B_x\hat{\underline{i}}+A_y\hat{\underline{ j}}\times
B_y\hat{\underline{ j}}+A_y\hat{\underline{ j}}\times
B_z\hat{\underline{k}}+\\&A_z\
hat{\underline{k}}\times B_x\hat{\underline{i}}+A_z\
$$\begin{aligned}\vec{\textbf{A}}\times\
vec{\textbf{B}}=&A_xB_x(\hat{\underline{i}}\times\
hat{\underline{i}})+A_xB_y(\hat{\underline{i}}\times\
hat{\underline{ j}})+A_xB_z(\hat{\underline{i}}\times\
hat{\underline{k}})+\\&A_yB_x(\hat{\underline{ j}}\ti
mes\
hat{\underline{i}})+A_yB_y(\hat{\underline{ j}}\times\
hat{\underline{ j}})+A_yB_z(\hat{\underline{ j}}\times\
hat{\underline{k}})+\\&A_zB_x(\hat{\underline{k}}\ti
mes\
hat{\underline{i}})+A_zB_y(\hat{\underline{k}}\times\
hat{\underline{ j}})+A_zB_z(\hat{\underline{k}}\times\
hat{\underline{k}})\end{aligned}$$
$$\begin{aligned}\vec A\times\vec
B=&A_xB_x(0)+A_xB_y(\underline{\hat{k}})+A_xB_z(-
\underline{\hat{ j}})+\\&A_yB_x(-
\underline{\hat{k}})+A_yB_y(0)+A_yB_z(\underline{\h
at{i}})+\\&A_zB_x(\underline{\hat{ j}})+A_zB_y(-
\underline{\hat{i}})+A_zB_z(0)\end{aligned}$$
$$\begin{aligned}\vec A\times\vec
70 | Statics
B=&A_yB_z(\underline{\hat{i}})+A_zB_y(-
\underline{\hat{i}})+\\&A_zB_x(\underline{\hat{ j}})+A
_xB_z(-
\underline{\hat{ j}})+\\&A_xB_y(\underline{\hat{k}})+
A_yB_x(-\underline{\hat{k}})\end{aligned}$$
$$ \begin{bmatrix}
\underline{\hat{i}} & \underline{\hat{ j}} &
\underline{\hat{k}} \\
A_x & A_y & A_z \\
B_x & B_y & B_z
\end{bmatrix} $$
For the first element of the first row, the i, take the
product down and to the right,
72 | Statics
( this yields iAyBz) minus the product down and to the
left
Key Takeaways
74 | Statics
vector that is a product of two vectors, perpindular
to the plane created from the two vectors.
1.6.1 Moments
76 | 1.6 Torque/Moment
Moments, like forces, can be represented as vectors
and have a magnitude, a direction, and a “point of
application”. For moments however a better name for
the point of application is the axis of rotation. This will
be the point or axis about which we will determine all
the moments.
Magnitude:
Direction:
1.6 Torque/Moment | 77
is a positive moment and a moment that would cause a
clockwise rotation is a negative moment.
78 | Statics
in the positive z direction and clockwise rotations are
represented by a vector in the negative z direction.
Axis of Rotation:
1.6 Torque/Moment | 79
or we will need to take the moments about the center of
mass of the body. Summing moments about other axes
of rotation will not result in valid calculations.
Calculating Moments:
80 | Statics
1.6.2 Scalar Method in 2 Dimensions
$$M=F\ast d$$
1.6 Torque/Moment | 81
moments while clockwise rotations are caused by
negative moments.
82 | Statics
Source: Engineering Mechanics, Jacob Moore et al.,
http://mechanicsmap.psu.edu/websites/
3_equilibrium_rigid_body/3-1_moment_scalar/
moment_scalar.html
1.6 Torque/Moment | 83
vector r from the point to anywhere on the line of
action of the force and the force vector itself.
84 | Statics
(r, F and M) are vectors. Before you can solve for the
cross product, you will need to write out r and F in
vector component form. Also, even for two dimensional
problems, you will need to write out all three
components of the r and F vectors. For two dimensional
problems the z components of the r and F vectors will
simply be zero, but those values are necessary for the
calculations.
The moment vector you get will line up with the axis
of rotation for the moment, where you can use the right
hand rule to determine if the moment is going clockwise
or counterclockwise about that axis.
1.6 Torque/Moment | 85
Finally, it is also important to note that cross product,
unlike multiplication, is not communicative. This means
that the order of the vectors matters, and r cross F will
not be the same as F cross r. It is important to always
use r cross F when calculating moments.
86 | Statics
Whether you use scalars: M = |r| |F| sinΘ or vectors: M = r x F
you can solve most moment/torque problem. The scalar method is
faster for 2-d problems, especially if the vectors are at 90 degree
angles from each other (sin 90º = 1). The vector method is more
robust, especially if there are additional angles involved. There is
the potential to make errors, so it’s recommended to use the Step
6 Review step to try multiple methods to ensure your answer is
correct.
See the examples in section 1.8 as many of them concern
moments.
Key Takeaways
1.6 Torque/Moment | 87
1.7 Problem Solving Process
Learning how to use a structured problem solving process will help
you to be more organized and support your future courses. Also, it
will train your brain how to approach problems. Just like basketball
players practice jump shots over and over to train their body how to
act in high pressure scenarios, if you are comfortable and familiar
with a structured problem solving process, when you’re in a high
pressure situation like a test, you can just jump into the problem like
muscle memory.
1. Problem
2. Draw
4. Approach
5. Analysis
◦ This is the actual solving step. This is where you show all
the work you have done to solve the problem.
◦ When you get an answer, restate the variable you are
solving for, include the unit, and put a box around the
answer.
6. Review
• It’s important to include the number and label the steps so it’s
clear what you’re doing, as shown in the example below.
• It’s okay if you make mistakes, just put a line through it and
keep going.
• Remember your header should include your name, the page
number, total number of pages, the course number, and the
assignment number. If a problem spans a number of pages, you
should include it in the header too.
90 | Statics
Key Takeaways
92 | Statics
1.8 Examples
Here are examples from Chapter 1 to help you understand these
concepts better. These were taken from the real world and supplied
by FSDE students in Summer 2021. If you’d like to submit your own
examples, please send them to:
[email protected]
1. Problem
1.8 Examples | 93
Source:
https://
www.fli
ckr.com
/
photos/
dejankr
smanov
ic/
3321820
7918
94 | Statics
2. Draw
Known:
x = 6 cm
y = 3.5 cm
Unknown: r, θ
4. Approach
5. Analysis
\begin{aligned}
&\tan \theta=\frac{y}{x} \\
&\tan \theta=\frac{3.5 \mathrm{~cm}}{6
\mathrm{~cm}} \\
&\theta=\tan ^{-1}\left(\frac{35}{6}\right) \\
&\theta=30.256^{\circ} \\
&\sin \theta=\frac{y}{r} \\
&r=\frac{y}{\sin \theta} \\
&r=\frac{35 \mathrm{~cm}}{\sin
\left(30256^{\circ}\right)} \\
&r=6946 \mathrm{~cm} \\
1.8 Examples | 95
&r=6.9 \mathrm{~cm}
\end{aligned}
6. Review
1. Problem
96 | Statics
between the fishing rod and fishing line is 45
degrees. If Mark catches a fish when 25 ft of
the fishing line is released while the fish is
diving down with a force of 180 N, how much
force does Mark need to apply (push down)
to the reel handle to bring in the fish? Draw
the position vector of the fish relative to the
reel.
Assumptions:
1.8 Examples | 97
Source: https://commons.wikimedia.org/wiki/
File:Deepsea.JPG
98 | Statics
2. Draw
Sketch:
Free-body diagram:
1.8 Examples | 99
Known:
rAB = 0.25 ft
rBC = 8 ft
rCD = 25 ft
FD = 180 N
θ = 45°
4. Approach
5. Analysis
[latex]\begin{aligned}&T_{D}=\left|r_{C D}\right| *
\left|F_{D}\right| * \sin \theta\\ &T_{D}=(7.62 m)(180
N) \sin \left(45^{\circ}\right)\\ &T_{D}=969.86766
\mathrm{Nm} \end{aligned}[/latex]
100 | Statics
Step 3: Solve for FA
[latex]\begin{aligned}&T_{A}=\left|r_{AB}\right| *
\left|F_{A}\right| * \sin \theta\\ &\text { Assume }
T_{A}=T_{D}\\
&F_{A}=\frac{T_{D}}{\left|r_{AB}\right| * \sin \theta}
\\ &F_{A}=\frac{969.86766 \mathrm{ Nm}}{0.0762
\mathrm{~m} \cdot \sin \left(45^{\circ}\right)} \\
&F_{A}=17,999.998 \mathrm{N} \\ &F_{A}=18,000
\mathrm{N} \end{aligned}[/latex]
Vector rAD:
6. Review
102 | Statics
Example 1.8.3: Dot product and cross
product, submitted by Anonymous ENGN
1230 Student
1. Problem
$$\underline{a}=[6\;\;\;5\;\;\;3]\;\;\;\u
nderline{b}=[8\;\;\;1\;\;\;3]$$
a) Find 6b
2. Draw
n/a
Known: a, b
4. Approach
5. Analysis
Part a:
$$6\underline{b}=6*[6\;\;\;5\;\;\;3]\\6\
underline{b}=[36\;\;\;30\;\;\;18]$$
Part b:
$$\underline{a}\cdot\
underline{b}=[6\;\;\;5\;\;\;3]\cdot[8\;\;\;1\;\;\;3]\
\=6\cdot 8+5\cdot 1+3\
cdot3\\=48+5+9\\\underline{a}\cdot\
underline{b}=62$$
Part c:
$$\underline{a}\times\underline{b}=\begin{bmatrix}
\underline{\hat{i}} &\underline{\hat{ j}} &
\underline{\hat{k}} \\
6 & 5 & 3 \\
8&1&3
\end{bmatrix}\\(5\cdot 3-3\cdot
1)\underline{\hat{i}}-(6\cdot 3-3\cdot
8)\underline{\hat{ j}}+(6\cdot 1-5\cdot
8)\underline{\hat{k}}\\\underline{a}\times\
underline{b}=12\underline{\hat{i}}+6\
underline{\hat{ j}}-34\underline{\hat{k}}$$
Part d:
$$ 2\
underline{a}=2*[6\;\;\;5\;\;\;3]=[12\;\;\;10\;\;\;6]\\
\underline{b}=[8\;\;\;1\;\;\;3]\\2\
underline{a}\times\underline{b} = \begin{bmatrix}
\underline{\hat{i}} & \underline{\hat{ j}} &
104 | Statics
\underline{\hat{k}} \\
12 & 10 & 6 \\
8&1&3
\end{bmatrix} \\=(10\cdot 3-6\cdot
1)\underline{\hat{i}}-(12\cdot 3-6\cdot
8)\underline{\hat{ j}}+(12\cdot 1-10\cdot
8)\underline{\hat{k}}\\2\underline{a}\times\
underline{b}=24\underline{\hat{i}}+12\
underline{\hat{ j}}-68\underline{\hat{k}}$$
6. Review
1. Problem
106 | Statics
Source: https://commons.wikimedia.org/wiki/
File:Girl_on_a_Bike_(Imagicity_116).jpg
2. Draw
108 | Statics
3. Knowns and Unknowns
Knowns:
Unknowns:
4. Approach
5. Analysis
Part a:
Part b:
[latex]T_2=\frac{T_1}{2}[/latex]
[latex]\frac{16 Nm}{2}=8 Nm[/latex]
110 | Statics
F_{A}\right| * \left| r \right|} \\ &\sin \theta=\frac{8
N m}{(100 N)(0.16 m)} \\ &\sin \theta=0.5 \\
&\theta=30^{\circ}, 150^{\circ}, etc
\end{aligned}[/latex]
Part c:
6. Review
1. Problem
112 | Statics
Assumptions: model the force as a single
point load acting on the door.
2. Draw
Sketch:
Free-body diagram:
Knowns:
• F = 100 N
• r1 = 45 cm
• r2 = 75 cm
• α = 45°
Unknowns:
4. Approach
$$ M=|r|\cdot|F|\cdot\sin\theta$$
5. Analysis
Part a)
114 | Statics
As shown below, the angle we find is also 45°. Now we
can continue and solve for M1 and M2.
$$\theta=90^{\circ}-45^{\circ}$$
$$\theta=45^{\circ}$$
$$ M_1=|r_1|\cdot|F|\cdot\sin\theta\\M_1=0.45m\
cdot 100N\
sin(45^{\circ})\\m_1=31.82Nm\\\\M_2=|r_2|\cdot
|F|\sin\theta\\M_2=0.75m\cdot 100N\cdot\
sin(45^{\circ})\\M_2=53.03Nm$$
Part b)
$$M=|r|\cdot|F|\cdot\sin\theta\\if \sin\theta=0,
M=0\\\sin\
theta=0\\\theta=\sin^{-1}(0)\\\theta=0^{\circ},
180^{\circ}, 360^{\circ}$$, etc
6. Review
1. Problem
116 | Statics
George to his dog?
b. What is Sparky’s velocity? (no need
to draw)
c. What is Sparky’s speed? (no need to
draw)
Source: https://www.piqsels.com/en/
public-domain-photo-oekac
2. Draw
Part a:
dg = [7 0 8] m Unknown: dsg= ?
ds = [0 6 6] m
Part b:
Part c:
4. Approach
118 | Statics
5. Analysis
Part a:
dsg= dg– ds
dsg =[7 0 8] m – [0 6 6] m
dsg =[7-0 0-6 8-6] m
dsg =[7 -6 2] m
Part b:
vsg=dsg/t
vsg=[7 -6 2] m/s
vsg=[7/4 -6/4 2/4] m/s
vsg=[1.75 -1.5 0.5] m/s
Part c:
vsg=vsg
vsg=(vsgx)2 + (vsgy)2 + (vsgz)2
vsg=(1.75)2 + (-1.5)2 + (0.5)2
vsg=2.36 m/s
6. Review
Part a:
Part b and c:
120 | Statics
CHAPTER 2: PARTICLES
In this chapter, we analyze our first static bodies (motion where
acceleration = 0), treating them as particles. The sections in this
chapter include:
• 2.1 Particle & Rigid Body – the difference between particles and
rigid bodies
• 2.2 Free Body Diagrams for Particles – learning how to model
forces and motion
• 2.3 Equilibrium Equations for Particles – analyzing static
bodies
• 2.4. Examples – problems submitted by other students.
Very simply, here are the important equations for this section (the
Ch 1 equations might be helpful too)
Particles are typically part of a larger scale, such as a sky diver falling
126 | Statics
through the sky, or a football flying through the air. Rigid body
analyses are required when the length or size of the object must be
considered, such as if you need to calculate the torque from turning
a bolt with a wrench, or if there is rotation, such as the bolt that is
being turned.
One way to think of it is that particles have mass, whereas rigid
bodies have mass and shape. We make an assumption that neither
particles nor rigid bodies deform (change shape). Note: we say
particles don’t deform even though we are already assuming that
the shape of particles is negligible.
In baseball, if you want to consider how far the ball travels, that
would be a particle analysis because the speed is much greater than
the size of the ball. A rigid body analysis would be how the bat
swings to hit the ball, because the length of the bat would change
how far the ball travels. A rigid body analysis could be to calculate
the spin on the ball as it flies through the air (if you focus on how it
is rotating).
You would have done particle analyses in your high school physics
classes. Starting in chapter 3, we’ll expand on these concepts to
include rigid bodies and bring shape and size into the problem.
Key Takeaways
128 | Statics
2.2 Free Body Diagrams for
Particles
A free-body diagram (FBD) helps you to simplify a complicated
problem. The first thing to remember is the object should always be
free which means, floating in space. You represent the floor or other
surfaces with forces. You might have done these particle free body
diagrams in your high school physics class, where all the forces act
at the centre of the object. (This will be different for rigid bodies).
To draw a free-body diagram remember four points:
For a baseball being hit by a bat (and neglecting air), the force of
gravity acts at the center, the force of the bat acts on the outside.
1
Notice in the figure the names FBat and FG are different in the figure
below. Also – you can understand what they represent quickly. Also
see the coordinate frame? You’ll be adding these in your sleep by the
end of this class.
130 | Statics
We are going to introduce the various kinds of forces
by means of examples. Here is the first
example:
132 | Statics
(1) The only thing touching the object while it
is up in the air (neglecting the air itself) is the
earth’s gravitational field. So there is only one
force on the object, namely the gravitational
force. The arrow representing the force vector is
drawn so that the tail of the arrow is touching
the object, and the arrow extends away from the
object in the direction of the force.
134 | Statics
To do this problem, you need the following
information about strings:
Key Takeaways
136 | Statics
2.3 Equilibrium Equations for
Particles
For a particle in static equilibrium, Newton’s 2nd law can be adapted
for [latex]\vec a = 0[/latex] and componentized in x y and z:
$$\sum\vec F=m*\vec a$$
$$\sum\vec F=0$$
$$\sum F_x=0\quad\quad\sum F_y=0\quad\quad\sum
F_z=0$$
Notice that the left size of the equation says ‘sum of the forces’
which means add up all the forces in that direction. In statics, they
will all cancel out. If you aren’t sure if something is in static motion,
sum the forces and see if they equal 0.
Static Equilibrium:
$$\sum\vec F=0$$
138 | Statics
names for any unknowns (either magnitudes or
directions).
Example:
$$F_g=(9.8)(6)\\F_g=58.8N\\\sum F_x=-T_1+T_2\
cos(15^{\circ})=0\\\sum F_y=T_2\
sin(15^{\circ})-58.8=0\\T_2=\frac{58.8}{\sin(15^{\cir
c})}=227.2N\\-T_1+227.2\
cos(15^{\circ})=0\\T_1=227.2\
140 | Statics
cos(15^{\circ})=219.4N\\T_1=219.4N\\T_2=227.2N$$
Key Takeaways
142 | Statics
2.4. Examples
Here are examples from Chapter 2 to help you understand these
concepts better. These were taken from the real world and supplied
by FSDE students in Summer 2021. If you’d like to submit your own
examples, please send them to the author [email protected].
No examples submitted from students, yet. In the mean time, here
are examples ( Example 1, Example 5, Example 8) from Engineering
Mechanics, Jacob Moore, et al. http://mechanicsmap.psu.edu/
websites/2_equilibrium_concurrent/
2-5_equilibrium_equations_particle/
equilibriumequationsparticle.html
▪ [latex]\underline{\hat{i}}\times\
underline{\hat{ j}}=\underline{\hat{k}}[/latex]
▪ [latex]\underline{\hat{ j}}\times\
underline{\hat{k}}=\underline{\hat{i}}[/latex]
▪ [latex]\underline{\hat{k}}\times\
underline{\hat{i}}=\underline{\hat{ j}}[/latex]
▪ [latex]\underline{\hat{ j}}\times\underline{\hat{i}}=-
\underline{\hat{k}}[/latex]
▪ [latex]\underline{\hat{k}}\times\underline{\hat{ j}}=-
\underline{\hat{i}}[/latex]
▪ [latex]\underline{\hat{i}}\times\underline{\hat{k}}=-
\underline{\hat{ j}}[/latex]
148 | Statics
that appears before the “×” in the mathematical
expression for the cross product; e.g. the A in A x B ).
150 | Statics
and envisioning how they are perpendicular to each other.
Try this one in 2d and 3d. Imagine (or draw) the right-angle
symbols (Answer will be in a few steps)
Example 1:
Using this x and y, let’s use the right-hand rule to find the direction
of z.
Sometimes you will need to flip your hand 180 degrees to find which
way lets you point your fingers in the y direction, for example:
152 | Statics
Example 3:
It’s important for you to be able to envision how the axes are
perpendicular. Now practice using the right hand rule if you are
trying to find x.
Keep going with these examples. The rules stay the same: thumb
towards z, curled fingers towards y, extended fingers towards x.
Find the missing axis:
.
.
Did you do it?
.
.
.
Here are the answers:
.
.
154 | Statics
the cross product r x F = M. The third method finds the scalar
value separately, then uses the right hand rule to find the direction
(positive or negative along the third axis).
The following will help you understand what is meant by: the
perpendicular part of position vector:
156 | Statics
vector r of the point of application of the force.
158 | Statics
From the diagram it is clear that the moment arm r is
just the magnitude of the component ┴ vector, in the
perpendicular-to-the-force direction, of the position
vector of the point of application of the force.
You use the right hand rule twice during this method to find the
vector. First to determine the coordinate frame and again to see
which the direction the torque is aligned. Then you multiply by the
magnitude of the perpendicular portion of the position vector (r⊥
or the “moment arm”) and the magnitude of the force vector. ):
Example 4:
If you find curling your fingers too confusing, you can try this
method that uses your thumb, pointer finger, and middle finger all
90 degrees apart. Your thumb is x, your pointer finger is y, your
middle finger is z.
160 | Statics
This is done by using your right hand, aligning your
thumb with the first vector and your index with the
second vector. The cross product will point in the
direction of your middle finger (when you hold your
middle finger perpendicular to the other two fingers).
This is illustrated in Figure A.14. Thus, you can often
avoid using equation A.1 and instead use the right hand
rule to determine the direction of the cross product and
equation A.2 to find its magnitude.
For axial vectors, you use what I’m calling the curly method. To find
whether the axis of rotation is positive or negative, curl your fingers
in the direction of rotation and your thumb shows the direction of
rotation, i.e. whether rotation is along the positive or negative x y
or z direction. (This assumes you already have a coordinate frame
defined to see which axis the wheel is rotating around and which
direction).
If a wheel is rolling, the axis is what it rolls around. Curl your
fingers in the direction of rotation and your thumb shows the
1
direction of rotation.
162 | Statics
Key Takeaways
164 | Statics
3.2 Couples
166 | Statics
Now we have some point A, which is distance x from
the first of the two forces. If we take the moment of
each force about point A, and then add these moments
together for the net moment about point A we are left
with the following formula.
$$M=-(F\ast x)+(F\ast(x+d))$$
Key Takeaways
168 | Statics
3.3 Distributed Loads
3.3.1 Intensity
170 | Statics
line, over a surface, or over a volume, that are
connected with a line or a surface as shown below.
An additional example:
172 | Statics
calculation, you get what you get. In this case we could
approximate this shape with two semi-circles on each
end of the wing with a triangle (∇"
role="presentation">∇) in the middle. For more accuracy
we could use a system similar to the trapezoidal rule.
174 | Statics
find the magnitude, direction, and point of application
of a single force that is equivalent to the distributed load
we are given. In this course we will only deal with
distributed loads with a uniform direction, in which case
the direction of the equivalent point load will match the
uniform direction of the distributed load. This leaves the
magnitude and the point of application to be found.
There are two options available to find these values:
$$F_{eq}=\int_{xmin}^{xmax}F(x)dx$$
176 | Statics
Now that we have the magnitude of the equivalent
point load such that it matches the magnitude of the
original force, we need to adjust the position (xeq) such
that it would cause the same moment as the original
distributed force. The moment of the distributed force
will be the integral of the force function (F(x)) times the
moment arm about the origin (x). The moment of the
equivalent point load will be equal to the magnitude of
the equivalent point load that we just found times the
moment arm for the equivalent point load (xeq). If we set
these two things equal to one another and then solve for
the position of the equivalent point load (xeq) we are left
with the following equation:
$$x_{eq}=\frac{\int_{xmin}^{xmax}(F(x)\ast
x)dx}{F_{eq}}$$
178 | Statics
The magnitude(Feq) of the equivalent point load will
be equal to the area under the force function. We can
find this area using calculus, but there are often easier
geometry based ways of finding the area under the force
function.
180 | Statics
Source: http://mechanicsmap.psu.edu/websites/
4_statically_equivalent_systems/4-5_equivalent_point_load
Example 4:
182 | Statics
Source: ” Equilibrium Structures, Support Reactions,
Determinacy and Stability of Beams and Frames” by
LibreTexts is licensed under CC BY-NC-ND .
https://eng.libretexts.org/Bookshelves/
Civil_Engineering/
Book%3A_Structural_Analysis_(Udoeyo)/01%3A_Chap
ters/
1.03%3A_Equilibrium_Structures_Support_Reactions_
Determinacy_and_Stability_of_Beams_and_Frames
For the following complex shape, this is how you find the
composite equivalent point force and location
([latex]\bar{x}[/latex]):
184 | Statics
Key Takeaways
We call the skywalk a cantilever beam and turn the real world
beam into a 2d model with constrains. So we can use the same
terminology, it is a fixed constraint, preventing horizontal
movement, vertical movement, and rotation.
2. Fixed
Notice that the Fixed restraint is the most restrictive and the
roller is the least restrictive. You put a force to show how the
restraint restricts motion. The roller only keeps the object from
moving vertically, so there is only 1 force. The pinned restraint
doesn’t allow horizontal or vertical movement, hence the two
forces. The fixed beam restricts vertical translation, horizontal
translation, and rotation, so there is a moment and two forces. Note
that this applies only to 2d restraints.
Here is a summary showing what motion is allowed by that type
of constraint:
188 | Statics
Typically reaction forces are either as follows: a pinned and a
fixed reaction force together (1 reaction force + 2 reaction forces =
3 restraints) or a fixed beam (2 reaction forces and 1 moment = 3
restraints).
190 | Statics
The characteristics of a rocker support are like those
of the roller support. Its idealized form is depicted in
Table 3.1.
3.4.4 Link
Example 1:
Example 3:
192 | Statics
Example 4:
194 | Statics
Key Takeaways
Once you have your equilibrium equations, you can solve them for
unknowns using algebra. The number of unknowns that you will be
able to solve for will be the number of equilibrium equations that
you have. In the x-y-z coordinate frame, there are 3 equations. so
there can be 3 unknowns. These are statically determinate.
Typically reaction forces are either as follows: a pinned and a
fixed reaction force together (1 reaction force + 2 reaction forces =
3 restraints) or a fixed beam (2 reaction forces and 1 moment = 3
restraints).
Indeterminate Loads
198 | Statics
Source: ” Equilibrium Structures, Support Reactions,
Determinacy and Stability of Beams and Frames” by
LibreTexts is licensed under CC BY-NC-ND .
https://eng.libretexts.org/Bookshelves/
Civil_Engineering/
Book%3A_Structural_Analysis_(Udoeyo)/01%3A_Chap
ters/
1.03%3A_Equilibrium_Structures_Support_Reactions_
Determinacy_and_Stability_of_Beams_and_Frames
Key Takeaways
200 | Statics
3.6 Examples
Here are examples from Chapter 3 to help you understand these
concepts better. These were taken from the real world and supplied
by FSDE students in Summer 2021. If you’d like to submit your own
examples, please send them to the author [email protected].
1. Problem
Source: https://www.maxpixel.net/
Seat-Couch-Interior-Home-Furniture-Room-Sofa-4
2817
2. Draw
Sketch:
202 | Statics
3. Knowns and Unknowns
Knowns:
• g = 9.81 m/s2
• mc = 30 kg
• mM = 60 kg
• mf = 70 kg
• Fg = 120 N
• rc = 1 m
• rM = 1.5 m
• rf = 4.5 m
• rB = 5 m
• rg = 2.5 m
Unknowns: NA, NB
4. Approach
Part a:
Part b:
204 | Statics
$$\sum F_y=0=N_A+N_B-F_C-F_M-F_f-
F_g\\\\N_A=F_C+F_M+F_f+F_g-
N_B\\\\N_A=(30kg\cdot 9.81m/s^2)+(60kg\cdot
9.81m/s^2)\\+(70kg\cdot 9.81m/
s^2)+120N-913.47N$$$$\\N_A=294.3N+588.6N+686.7N+
120N-913.47N$$$$\\N_A=776.13N\\\\\\underline{N_
A=776N}$$
6. Review
1. Problem
Real-life scenario:
Source: https://www.pxfuel.com/en/
free-photo-ekahu
206 | Statics
2. Draw
Sketch:
Free-body diagram:
Known:
d = 10 in
F = 7.5 lb
Unknown: M
4. Approach
5. Analysis
Find radius:
$$r=\frac{d}{2}\\r=\frac{10in}{2}\\r=5in\\5in\
times\frac{1ft}{12in}=0.42ft$$
$$ \underline{r}_A= \begin{bmatrix}
0.42 \\
0
\end{bmatrix}ft\:\; \underline{F}_A=\begin{bmatrix}
0 \\
7.5
\end{bmatrix}lb \\\underline{r}_B=\begin{bmatrix}
-0.42 \\
0
\end{bmatrix}ft\:\; \underline{F}_B=\begin{bmatrix}
0 \\
-7.5
\end{bmatrix}lb $$
208 | Statics
Find MA:
$$\underline{M}_A=\underline{r}_A\times
\underline{F}_A=\begin{bmatrix}
\underline{\hat{i}} & \underline{\hat{ j}} &
\underline{\hat{k}} \\
0.42 & 0 & 0 \\
0 & 7.5 & 0
\end{bmatrix}$$ $$\underline{M}_A=\hat{i}
\begin{bmatrix}
0 & 0 \\
7.5 & 0
\end{bmatrix} -\underline{\hat{ j}} \begin{bmatrix}
0.42 & 0 \\
0&0
\end{bmatrix}+\underline{\hat{k}} \begin{bmatrix}
0.42 & 0 \\
0 & 7.5
\end{bmatrix}\\\underline{M}_A=(\underline{\hat{i}}(
0)-\underline{\hat{ j}}(0)+\underline{\hat{k}}(o.42\cdot
7.5-0\cdot 0))ft\cdot lb\\\underline{M}_A=3.15\
underline{\hat{k}} ft\cdot lb$$
Find MB:
$$\underline{M}_B=\underline{r}_B\times
\underline{F}_B=\begin{bmatrix}
\underline{\hat{i}} & \underline{\hat{ j}} &
\underline{\hat{k}} \\
-0.42 & 0 & 0 \\
0 & -7.5 & 0
\end{bmatrix}$$$$\underline{M}_B=\hat{i}
\begin{bmatrix}
0 & 0 \\
$$\underline{M}=\underline{M}_A+\underline{M}_B
\\\underline{M}=3.15ft\cdot lb+3.15ft\cdot
lb\\\underline{M}=6.3\underline{\hat{k}}ft\cdot lb$$
6. Review
$$M=f\cdot d\\M=7.5lb\cdot
10in(\frac{1ft}{12in})\\M=7.5lb\cdot
\frac{5}{6}ft\\\\M=6.25ft\cdot lb$$
210 | Statics
Example 3.6.3: Distributed Load, Submitted
by Luciana Davila
1. Problem
w = 4x4 +2 N/m.
Sketch:
Free-body diagram:
212 | Statics
3. Knowns and Unknowns
Knowns:
• w = 4x4 + 2 N/m
• L = 100 cm
• xmin = 0
• xmax = 1
Unknowns: xr, Fr
4. Approach
$$F_r=\int^{xmax}_{xmin}
wdx\\X_r=\frac{\int^{xmax}_{xmin}
x*w(x)*dx}{\int^{xmax}_{xmin}wdx}$$
5. Analysis
$$ F_r=\int^1_0 (4x^4+2)dx\;\;
N\\F_r=(\frac{4x^5}{5}+2x)\vert^1_0\;\;N\\F_r=(\f
rac{4}{5}+2)N\\F_r=2.8N$$
$$X_r=\frac{(\int^1_0x(4x^4+2)dx)N/
m}{\int^1_0(4x^4+2)dx)N}\\X_r=\frac{\int^1_0(4x^
5+2x)dxN/
m}{2.8N}\\X_r=\frac{(\frac{4x^6}{6}+\frac{2x^2}{2})\
vert^1_0N/m}{2.8N}\\X_r=\frac{(\frac{2}{3}+1)N/
m}{2.8N}\\X_r=0.59m$$
6. Review
214 | Statics
CHAPTER 4: RIGID BODIES
This is arguably the most fundamental chapter for Statics. Learn
these concepts and the next two chapters will make a lot of sense.
Without this chapter, the next chapters will be much more
confusing. When people talk about Statics, this chapter contains the
concepts they are talking about. You will use free-body diagrams
and the equilibrium equations in many other courses. Here are the
sections in this Chapter:
• Gravity: Fg = mg
• Normal: Calculated
• Friction: Ff = mN
• Spring: FS = -kx
• Applied: Measured or calculated
Normal Force
218 | Statics
stationary on the table. The table sags quickly and the
sag is slight, so we do not notice it. But it is similar to
the sagging of a trampoline when you climb onto it.
$$N=mg$$
220 | Statics
on the object is divided into two components: a force
acting perpendicular to the plane, wywy"
role="presentation" style="font-family: proxima-nova,
sans-serif;padding: 1px 0px;margin: 0px;font-size:
17.44px;vertical-align: baseline;background:
#ffffff;border: 0px;line-height: 0;text-indent: 0px;text-
align: left;text-transform: none;font-style: normal;font-
weight: 400;letter-spacing: normal;float: none;direction:
ltr;max-width: none;max-height: none;min-width:
0px;min-height: 0px;color: #373d3f">, and a force
acting parallel to the plane, wx. The normal force
[latex]\vec N[/latex] N→." role="presentation"
style="font-family: proxima-nova, sans-serif;padding:
1px 0px;margin: 0px;font-size: 17.44px;vertical-align:
baseline;background: #ffffff;border: 0px;line-height:
0;text-indent: 0px;text-align: left;text-transform:
none;font-style: normal;font-weight: 400;letter-spacing:
normal;float: none;direction: ltr;max-width: none;max-
height: none;min-width: 0px;min-height: 0px;color:
$$w_x=w\sin\theta=mg\sin\theta$$
and
$$w_y=w\cos\theta=mg\cos\theta$$
$$N=mg\cos\theta$$
222 | Statics
vectors. The angle θ θ" role="presentation" style="font-
family: proxima-nova, sans-serif;padding: 1px
0px;margin: 0px;font-size: 17.44px;vertical-align:
baseline;background: #ffffff;border: 0px;line-height:
0;text-indent: 0px;text-align: left;text-transform:
none;font-style: normal;font-weight: 400;letter-spacing:
normal;float: none;direction: ltr;max-width: none;max-
height: none;min-width: 0px;min-height: 0px;color:
#373d3f">of the incline is the same as the angle formed
between w and wy. Knowing this property, we can use
trigonometry to determine the magnitude of the weight
components:
$$\cos\theta=\frac{w_y}{w},\:w_y=w\cos\
theta=mg\cos\theta\\\sin\
theta=\frac{w_z}{w},\:w_x=w\sin\theta=mg\sin\
theta$$
Tension
$$F_{net}=T-w=0$$
where T and w are the magnitudes of the tension and
weight, respectively, and their signs indicate direction,
with up being positive. As we proved using Newton’s
second law, the tension equals the weight of the
supported mass:
$$T=w=mg$$
Thus, for a 5.00-kg mass (neglecting the mass of the
rope), we see that
$$T=mg=(5.00kg)(9.80m/s^2)=49.0N$$
224 | Statics
4.1 External Forces | 225
Flexible connectors are often used to transmit forces
around corners, such as in a hospital traction system, a
tendon, or a bicycle brake cable. If there is no friction,
the tension transmission is undiminished; only its
direction changes, and it is always parallel to the flexible
connector, as shown below:
$$T=\frac{w}{2\sin\theta}$$
$$T=\frac{F\perp}{2\sin\theta}$$
226 | Statics
horizontal (i.e., θ=0 and sin θ=0). For example, the image
below shows a situation where we wish to pull a car out
of the mud when no tow truck is available. Each time the
car moves forward, the chain is tightened to keep it as
straight as possible. The tension in the chain is given
by[latex]T=\frac{F\perp}{2\sin\theta}[/latex] and
since θ is small, T is large. This situation is analogous to
the tightrope walker, except that the tensions shown
here are those transmitted to the car and the tree
rather than those acting at the point where F⊥ is
applied.
Friction
228 | Statics
Source: University Physics Volume 1, Openstax CNX.
https://courses.lumenlearning.com/suny-
osuniversityphysics/chapter/5-6-common-forces/
230 | Statics
4.2 Rigid Body Free Body
Diagrams
Following what we learned in Section 2.2 on particle Free-Body
Diagrams (FBDs), this section will expand on that for rigid bodies.
The biggest difference between a particle and rigid body FBD is
where the force is applied. In a rigid body FBD, you have to be
precise about pointing the head of the force arrow to the location
where it applied. For example. if we wanted to make a FBD of you
and me high-5’ing, you would apply the force from your hand onto
my hand, not at my center of mass.
In this section, first we will learn how to do a FBD for a part, then
we look at how to model a system of multiple objects.
Here are some tips to keep in mind about each of the forces:
1. Draw shape
2. Add coordinate frame
3. Replace forces with arrows
4. Label each force uniquely
To model a book being pushed across the table, you would apply the
following forces at the following locations (see image below)
232 | Statics
• the frictional force running along the bottom surface between
the book and table (yellow arrow)
• the gravitational force acting at the center of mass (pink arrow)
• any applied force at the point of application, such as your hand
pushing on the book (blue arrow)
If instead, the book were being pulled by a string, the image would
be the same but the applied force and frictional force would change
direction (because friction always opposes motion).
234 | Statics
The first step in solving most mechanics problems will
be to construct a free body diagram. This simplified
diagram will allow us to more easily write out the
equilibrium equations for statics or strengths of
materials problems, or the equations of motion for
dynamics problems.
236 | Statics
another object or surface will
experience a normal force that
is perpendicular (hence normal)
to the surfaces in contact.
238 | Statics
cannot be used for pushing.
240 | Statics
◦ The system should be floating in space with no surface
(such as the floor)
◦ Include coordinate frame
◦ Use only external forces on system FBD (gravity, applied,
normal, frictional, spring).
◦ DO NOT include internal forces
◦ It is especially important to use unique labels, so the top
book forces are labelled 1, and the bottom book forces are
labelled 2 (or T for top and B for bottom, or A and B).
2. Draw a FBD for each part separately & coord frame with equal
and opposite arrows for internal system forces
Some tips:
4.2.3 Examples
242 | Statics
brakes locking up all four wheels. The distance between
the two wheels is 8 feet and the center of mass is 3 feet
behind and 2.5 feet above the point of contact between
the front wheel and the ground. Draw a free body
diagram of the car as it comes to a stop.
244 | Statics
Example 2: Part FBD (a beam)
246 | Statics
the cart and barrels (C and A/B) in red and between the barrels
(A & B) in blue. Notice the matching labels for internal forces but
opposing directions. Notice that the coordinate frame has been
rotated consistently in all of the FBDs.
Key Takeaways
248 | Statics
4.3 Rigid Body Equilibrium
Equations
We use the equilibrium equations to calculate any unknown forces
& moments using the known forces and values, and the following
equations:
The particle equilibrium equations were covered in section 2.3.
These are:
$$
\Sigma F_{x}=0, \Sigma F_{y}=0, \Sigma F_{z}=0
$$
Now for a rigid body where forces are analyzed at different points
on a body, we can take moments into account. There are 3 equations
for 2d and 4 equations for 3d:
Rigid Body-Two Dimensions
$$
\Sigma F_{x}=0, \Sigma F_{y}=0, \Sigma M_{O}=0
$$
Rigid Body-Three Dimensions
$$
\begin{gathered}
\Sigma F_{x}=0, \Sigma F_{y}=0, \Sigma F_{z}=0 \\
\Sigma M_{x^{\prime}}=0, \Sigma M_{y^{\prime}}=0, \Sigma
M_{z^{\prime}}=0
\end{gathered}
$$
Because these are static bodies, the right side of the equations
equal 0. In dynamics, they will equal the mass times the acceleration
for translation and rotation.
250 | Statics
where the sum of the components in the x, y, and z
direction must be equal to zero. The body may also have
moments about each of the three axes. The second set
of three equilibrium equations states that the sum of the
moment components about the x, y, and z axes must
also be equal to zero.
252 | Statics
Source: Engineering Mechanics, Jacob Moore, et al.
http://mechanicsmap.psu.edu/websites/
3_equilibrium_rigid_body/
3-6_equilibrium_equations_rigid_body/
equilibrium_equations_rigid_body.html
Example 1:
The car below has a mass of 1500 lbs with the center
of mass 4 ft behind the front wheels of the car. What are
the normal forces on the front and the back wheels of
the car?
254 | Statics
Source: Engineering Mechanics, Jacob Moore et al.,
http://mechanicsmap.psu.edu/websites/
3_equilibrium_rigid_body/
3-6_equilibrium_equations_rigid_body/pdf/
EquilibriumEquationsExtended_WorkedProblem1.pdf
256 | Statics
Source: Engineering Mechanics, Jacob Moore et al.,
http://mechanicsmap.psu.edu/websites/
3_equilibrium_rigid_body/
3-6_equilibrium_equations_rigid_body/pdf/
EquilibriumEquationsExtended_WorkedProblem5.pdf
258 | Statics
4.4 Friction and Impending
Motion
Dry Friction
260 | Statics
friction, the magnitude of the friction force opposing
motion will be equal to the kinetic coefficient of friction
times the normal force between the box and the
surface. The kinetic coefficient of friction also depends
upon the two materials in contact, but will almost
always be less than the static coefficient of friction.
262 | Statics
Determining the Force Required to Make an Object
“Tip”:
264 | Statics
Source: Engineering Mechanics, Jacob Moore, et al.
http://mechanicsmap.psu.edu/websites/7_friction/
7-2_slipping_vs_tipping/slippingvstipping.html
Example 1
Example 2:
266 | Statics
4.4 Friction and Impending Motion | 267
Source: Engineering Mechanics, Jacob Moore, et al.
http://mechanicsmap.psu.edu/websites/7_friction/
7-1_dry_friction/pdf/
DryFriction_WorkedExample1.pdf
Example 3:
268 | Statics
degrees. If the static coefficient of friction between the
ice and the sled is .4 and the kinetic coefficient of
friction is .3, what is the required pulling force needed
to keep the sled moving at a constant rate?
270 | Statics
Example 4:
274 | Statics
Source: Engineering Mechanics, Jacob Moore, et al.
http://mechanicsmap.psu.edu/websites/7_friction/
7-2_slipping_vs_tipping/pdf/
TippingVsSlipping_WorkedExample1.pdf
Key Takeaways
276 | Statics
4.5 Examples
Here are examples from Chapter 4 to help you understand these
concepts better. These were taken from the real world and supplied
by FSDE students in Summer 2021. If you’d like to submit your own
examples, please send them to the author [email protected].
1. Problem
Real-life scenario:
2.Draw
Sketch:
278 | Statics
Free-body diagram:
Known:
• rBi = 2 ft
• rJ = 9 ft
• rB = 11 ft
• FBi = 160 lb
• FJ = 145 lb
Unknown: rBo
4. Approach
5. Analysis
$$ \sum F_y=0=-F_{Bi}-F_{Bo}-
F_J+A_y+B_y\\\\B_y=F_{Bi}+F_{Bo}+F_J-
A_y\\\\B_y=160 lb+180 lb+145 lb-225
lb\\\\\\B_y=260 lb$$
$$\sum M_A=0=-(F_{Bi})(r_{Bi})-(F_{Bo})(r_{Bo})-
(F_{J})(r_{J})+(B_{y})(r_{B})\\r_{bo}=\frac{-
(F_{Bi})(r_{Bi})-
(F_{J})(r_{J})+(B_{y})(r_{B})}{F_{Bo}}\\r_{Bo}=\frac{-
(160 lb)(2 ft)-(145 lb)(9 ft)+(260 lb)(11 ft)}{180 lb}$$
$$\underline{r_{Bo}=6.86 ft}$$
6. Review
280 | Statics
with no problems. If Bobby were to stand between 0 ft
and 6.811 ft, the left side of the bridge would fail.
1. Problem
θ = 15°
FA
4. Approach
282 | Statics
Draw the box, then draw all forces acting on it
5. Analysis
6. Review
1. Problem
2. Draw
Sketch:
284 | Statics
Free Body Diagram:
Knowns:
• FA = 150 N
• θ = 30°
• m = 12 kg
Unknowns: FN, μ
5. Analysis
Part a:
Find Fg:
$$Fg=m\cdot g\\Fg=(12kg)(9.81m/
s^2)\\\\Fg=117.72N$$
$$\sum Fy=0=F_N-F_g-F_A\sin
30^{\circ}\\0=F_N-117.72N-150N\cdot \sin
30^{\circ}\\F_N=117.72+150N\cdot\sin
30^{\circ}\\\\\underline{F_N=192.7 N}$$
Part b:
$$F_f=150N\cdot\cos 30^{\circ}\\F_f=M\cdot
F_N\\150N\cdot\cos 30^{\circ}=\mu\cdot F_N$$
$$\mu=\frac{150N\cdot\cos
30^{\circ}}{F_N}\\\mu=\frac{150N\cdot\cos
30^{\circ}}{192.72N}=0.67405$$
$$\underline{\mu=0.67}$$
6. Review
286 | Statics
FN, Fg, and the y component of FA are the only forces
in the y direction so it makes sense that they need to
equal zero for the equilibrium equations. FN is the only
positive force in the y direction, so it makes sense that it
equals the magnitude of the other two put together.
1. Problem
288 | Statics
2. Draw
Sketch:
Knowns:
Unknowns:
4. Approach
5. Analysis
Part a:
$$\sum F_y=0=-
F_g+F_f\\F_f=F_g\\F_f=49.05N$$
$$F_f=\mu_1
F_N\\F_N=\frac{F_f}{\mu_1}\\F_N=\frac{49.05N}{0
.49}$$
$$\underline{F_N=100.1N}$$
Part b:
$$\sum F_x=0=F_N-
F_A\\F_A=F_N\\F_A=100.1N$$
290 | Statics
$$w=\frac{F}{L}\\w={100.1N}{(16cm\times\
frac{1m}{100cm})}$$
$$\underline{w=625N/m}$$
6. Review
1. Problem
2. Draw
Sketch:
292 | Statics
3. Knowns and Unknowns
• θA = 20°
• θB = 60°
• θC = 15°
• m = 20 kg
• FA = 200 N
• FB = 150 N
Unknowns: Ff
4. Approach
5. Analysis
Part a:
Step 1: find FG
294 | Statics
$$F_{GX}=F_G\sin(15^{\circ})\\F_{GX}=196.2N\
cdot\sin(15^{\circ})\\F_{GX}=50.78N$$
$$\cos(20^{\circ})=\frac{F_A}{F_{AX}}\\F_{AX}=\fr
ac{F_A}{\cos(20^{\circ})}\\F_{AX}=\frac{200N}{\cos(
20^{\circ})}\\F_{AX}=212.8N$$
$$F_{BX}=F_B\cos(60^{\circ})\\F_{BX}=150N\
cos(60^{\circ})\\F_{BX}=75N$$
$$\underline{F_f=-87.02N}$$
6. Review
296 | Statics
CHAPTER 5: TRUSSES
This chapter will introduce you to a special type of structure called
a ‘truss’. You’ll analyze these structures more in your Structures
course, but for Statics you will need to know how to calculate the
force in each member, using two methods: method of joints and
method of sections. At first this might seem confusing, but there is
something quite elegant and magical about the method once you
understand it. Here are the sections in this Chapter:
300 | Statics
only forces, no moments) acting on it in only two
locations. In order to have a two force member in static
equilibrium, the net force at each location must be
equal, opposite, and collinear. This will result in all two
force members being in either tension or compression
as shown in the diagram below.
302 | Statics
5.1.2 Trusses
A truss is an
engineering structure
that is made entirely
of two force members. In
addition, statically
determinate trusses
(trusses that can be
analyzed completely
Adapted from image by
using the equilibrium ToddC4176 CC-BY-SA 3.0
equations), must
be independently rigid. This means that if the truss was
separated from its connection points, no one part would
be able to move independently with respect to the rest
of the truss.
304 | Statics
Source: https://eng.libretexts.org/Bookshelves/Civil_Engineering/
Book%3A_Structural_Analysis_(Udoeyo)/01%3A_Chapters/
1.05%3A_Internal_Forces_in_Plane_Trusses
Source: https://eng.libretexts.org/Bookshelves/Civil_Engineering/
Book%3A_Structural_Analysis_(Udoeyo)/01%3A_Chapters/
1.05%3A_Internal_Forces_in_Plane_Trusses
• joints
• members, and
• external forces (reaction forces and applied forces).
The joints are often labelled with a letter and are where the
external forces and members connect.
306 | Statics
Source: Engineering Mechanics, Jacob Moore, et al.
http://mechanicsmap.psu.edu/websites/5_structures/
5-4_method_of_joints/methodofjoints.html
The members are the metal or wooden beams that are labelled
with the connection between joints. For example member AB
connects joints A and B.
The applied force / load from trucks and cars goes from the deck,
to the stringers, across the beams, to the joints of the truss where it
is carried as applied (external) forces on the edges of the bridge.
308 | Statics
Image annotated from original source: https://upload.wikimedia.org/
wikipedia/en/2/25/Nine_stringers%2C_2_floorbeams.jpg
.
.
.
.
.
.
Any ideas?
.
.
.
.
.
Here’s the answer!
310 | Statics
Annotations added from original source: https://www.maxpixel.net/static/
photo/1x/Buildings-Leaves-Park-Autumn-Road-Fall-Structure-5623840.jpg
312 | Statics
Source: https://www.flickr.com/photos/121935927@N06/13580545445
Key Takeaways
314 | Statics
Basically: A truss is a rigid structure composed of two
force members (where forces are applied at only two
locations) that connect at joints and have external forces
applied. The internal forces of the truss put members in
compression (-) or tension (+).
Now solve for the reaction forces (Rax Ray Re) looking only at the
external forces using the equilibrium equations for a rigid body:
$$\sum F_x=0\\\sum F_y=0\\\sum M=0$$
Assuming the length of each member is L:
$$\sum F_x=R_{ax} = 0, \\\underline{R_{ax} = 0}$$
Next, pick a joint where there are 2 or fewer unknown values such
as a or e. This is because you only have 2 equations available to find
the unknowns: [latex]\sum F_x=0 \text{, } \sum F_y=0[/latex].
The following table shows the number of known and unknown
forces at each joint.
Joint: a b c d e f g
Known
2 0 0 0 1 1 1
forces:
Unknown
2 3 4 3 2 4 4
forces:
Member ab bc cd de ef fg ag bg cg cf df
Force (lb) 96.2 96.2 77.0 77.0 38.5 86.6 48.1 96.2 19.3 19.3 77.0
Tension or
C C C C T T T T T C T
Compression
And that’s it! If you don’t specify compression or tension, you should
use positive and negative to denote tension and compression,
respectively.
318 | Statics
The method of joints is a process used to solve for the
unknown forces acting on members of a truss. The
method centers on the joints or connection points
between the members, and it is usually the fastest and
easiest way to solve for all the unknown forces in a truss
structure.
320 | Statics
to assume all forces are tensile, then later in
the solution any positive forces will be
tensile forces and any negative forces will be
compressive forces.
◦ Label each force in the diagram. Include
any known magnitudes and directions and
provide variable names for each unknown.
Example 1:
322 | Statics
Solution:
Find the force acting in each of the members of the truss shown
below. Remember to specify if each member is in tension or
compression.
Solution here.
In summary:
326 | Statics
Key Takeaways
328 | Statics
5.3 Method of Sections
The method of sections uses rigid body analysis to solve for a specific
member or two. Instead of looking at each joint, you make a cut
through the truss, turning the members along that line into internal
forces (assume in tension). Then you solve the rigid body using the
equilibrium equations for a rigid body: [latex]\sum F_x=0\;\sum
F_y=0\;\sum M_z=0[/latex]
The truss:
For this example, you could choose the right half or left half. For
some problems, being strategic is necessary otherwise you need to
make multiple cuts. In this problem you had to solve for the reaction
forces first, but that isn’t always the case as you can sometimes just
make the cut (see example 2 below).
Here are more examples of how to make a cut and showing the
naming convention:
330 | Statics
Source: Internal Forces in Beams and Frames,
Libretexts. https://eng.libretexts.org/Bookshelves/
Civil_Engineering/
Book%3A_Structural_Analysis_(Udoeyo)/01%3A_Chap
ters/1.05%3A_Internal_Forces_in_Plane_Trusses
332 | Statics
2. Treating the entire truss structure as a rigid
body, draw a free body diagram, write out the
equilibrium equations, and solve for the external
reacting forces acting on the truss structure. This
analysis should not differ from the analysis of a
single rigid body.
334 | Statics
any known magnitudes and directions and
provide variable names for each unknown.
Example 1:
336 | Statics
5.3 Method of Sections | 337
338 | Statics
Source: Engineering Mechanics, Jacob Moore, et al.
http://mechanicsmap.psu.edu/websites/5_structures/
5-5_method_of_sections/pdf/
MethodOfSections_WorkedExample1.pdf
Example 2:
340 | Statics
5.3 Method of Sections | 341
Source: Engineering Mechanics, Jacob Moore, et al.
http://mechanicsmap.psu.edu/websites/5_structures/
5-5_method_of_sections/pdf/
MethodOfSections_WorkedExample2.pdf
In summary:
342 | Statics
Key Takeaways
344 | Statics
5.4 Zero-Force Members
This is a special case that is specially useful for method of joints and
method of sections. These special types of members called zero-
force members ensure the truss stays in a particular shape as a rigid
body, but carries no load.
Zero-force members are members that you can tell just by
inspection that they carry no load. They are important to the
structure to ensure it stays in a rigid shape.
Zero-force members can be found considering the equilibrium
equations. Look at joint e below. In the y direction, there is only
1 force: Feh. So if the sum of the forces in the y direction $latex
\sum F_{eh} = Feh = 0, then Feh = 0. Similarly, Fmk and Fcp are zero-
force members (if you look at joint m and c). Note that if you looked
at joint k or p, you couldn’t tell that Fmk and Fcp are zero-force
members.
There isn’t a huge problem if you can’t find zero-force members just
from inspection, but you might find that certain joints are not able
to be solved as easily. (Zero-force members let you have one less
unknown).
346 | Statics
Also see that L and G have no zero-force members because the
externally applied loads balance the members.
Here are some examples to practice on:
Example 1
Source: https://commons.wikimedia.org/wiki/File:Camelback-truss.svg
Example 2
Source: https://commons.wikimedia.org/wiki/File:Bowstring-truss.svg
Example 3
Source: https://pxhere.com/en/photo/995729
348 | Statics
Key Takeaways
Application: In trusses.
1. Problem
Source: https://flic.kr/p/txjSpP
Sketch:
352 | Statics
Free-body Diagram:
Known:
• P = 500 N
4. Approach
5. Analysis
Part a:
$$\sum F_x=0=P-
R_{Ax}\\R_{Ax}=P\\R_{Ax}=500N$$
$$\sum F_y=0=R_{By}+R_{Ay}\\R_{Ay}=-
R_{By}\\R_{Ay}=-250N$$
354 | Statics
frame). This makes sense as RAy and RBy are the only
external forces in the y direction, so they have to cancel
each other for the equilibrium equations to be true.
Therefore, one of them should have a negative direction.
We will leave this answer as is for now, but the next time
we draw the system, we will change the direction of the
arrow.
$$ \underline{R_{Ax}=500N,\; R_{Ay}=-250N, \;
R_{By}=250N}$$
Part b:
$$\sum
F_x=0=P+\frac{4}{\sqrt{17}}F_{CG}\\\frac{4}{\sqrt{17}}
F_{CG}=-P\\F_{CG}=-
P(\frac{\sqrt{17}}{4})\\F_{CG}=-500N(\frac{\sqrt{17}}{4
})\\F_{CG}=-515.388N$$
Part c:
356 | Statics
quickly solved using equilibrium equations. Had the
question asked for all member loads to be solved
however, the method of joints would have been the
faster approach.
6. Review
Part a:
Part b:
Part c:
1. Problem
358 | Statics
2. Draw
Free-body diagram:
4. Approach
5. Analysis
Part a:
360 | Statics
If we do the same type of analysis for the other joints
and remove the zero-force members, the structure now
looks like this:
Part b:
6. Review
362 | Statics
CHAPTER 6: INTERNAL
FORCES
In the last chapter we looked at the normal (axial) force running
through beams joined into trusses by analyzing either the joints or
a whole section of the truss.
In this chapter, we look at what happens along a single beam. We
will look at three types of internal forces and moments. Note that
when we say ‘internal forces’, we really mean ‘internal forces and
moments’. Inside a beam, we will calculate the normal and shear
forces as well as the bending moment at any point in the beam.
For this chapter: the shear force and bending moment change
throughout the beam because additional transverse forces are
applied. However, the normal force usually stays the same, because
it’s uncommon to have applied axial forces along the beam.
Here are the sections in this Chapter:
Direction
Force/ for a
Abbreviation Unit
Moment horizontal
beam
Normal
N N or lb horizontal
Force
Nm or
Moment M rotation
ft-lb
Note that for a vertical column, the normal force would be vertical.
For this reason, the normal force is often called ‘axial’ as in: along
the axis. The shear force for a column would be horizontal and is
sometimes called ‘transverse’.
This is for a 2d analysis of the beam assuming there is negligible
loading in the third dimension.
366 | Statics
force, and the bending moment, as shown in section k of
the cantilever of the figure below. To predict the
behavior of structures, the magnitudes of these forces
must be known. In this chapter, the student will learn
how to determine the magnitude of the shearing force
and bending moment at any section of a beam or frame
and how to present the computed values in a graphical
form, which is referred to as the “shearing force” and
the “bending moment diagrams.” Bending moment and
shearing force diagrams aid immeasurably during
design, as they show the maximum bending moments
and shearing forces needed for sizing structural
members.
Normal Force
Shearing Force
Bending Moment
368 | Statics
• 1 normal force (N)
• 2 shear forces (V1 & V2), and
• 3 bending moments (M1, M2, & T – torsion).
When you look at the beam as a whole (in the figure below), positive
shear is right side down. When you cut into beam, for it to be in
static equilibrium, the positive shear must then be up on the right
to be equal and opposite of the overall motion.
370 | Statics
Axial (Normal) Force
Shear Force
Bending Moment
372 | Statics
forces? b) what are the internal forces at the midpoint b) between
reaction forces?
[latex]\sum F_{X}=A_{x}=0[/latex]
[latex]\sum F_{y}=A_{y}+C-\omega L=0[/latex]
[latex]\sum M_{A}=-(\omega L)\left(\frac{L}{2}\right)+d_{A B}
C=0[/latex]
2. Make a cut at B.
3. In a FBD of one side of the cut, add the internal forces (and
moments) using the positive sign convention.
374 | Statics
$$ \sum M_A = – (w * 2 ft) * (1 ft) – V * (2 ft) + M = 0 \\ M = (100
\frac{lb}{ft}) * 2 ft^2 + (-112.5 lb) * (2 ft) \\ M = 200 ft \cdot lb –
225 ft \cdot lb \\ M = -25 ft \cdot lb \text{ (- indicates going
reverse direction)} $$
Key Takeaways
378 | Statics
at a section x + dx be V + dV and M + dM, respectively.
The total load acting through the center of the
infinitesimal length is wdx.
[latex]M+dM=M+Vdx[/latex]
or
or
[latex]\frac{d^2M}{dx^2}=-w(x)[/latex] (Equation
6.5)
380 | Statics
6.2.3 Producing a Shear/Moment Diagram
Example 1
382 | Statics
at B is fixed, there will be three reactions at that support, namely
By, Bx, and MB. Applying the conditions of equilibrium suggests the
following:
[latex]\sum F_{x}=0: \quad \underline{B_{x}=0}[/latex]
[latex]\sum F_{y}=0: \quad-5 lb+B_{y}=0[/latex]
[latex]\qquad \quad \underline{B_{y}=5 lb}[/latex]
[latex]\sum M_{B}=0: \quad(5 lb )(3 \mathrm{ft})-M=0[/latex]
[latex]\qquad \quad \underline{M=15 ft \cdot lb}[/latex]
3. Make a cut and add internal forces N V and M using the positive
sign convention. Depending on the number of loads, you may need
multiple cuts
Only 1 cut needed because only 1 load is added at the end. (If it
were in the middle there would be 2 sections to consider). The value
x could be 0 to 3 ft.
384 | Statics
The obtained expression is valid for the entire beam (the region
0 < x < 3 ft). The negative sign indicates a negative moment, which
was established from the sign convention for the moment, so the
moment actually goes in the opposite direction. The moment due
to the 5 lb force tends to cause the segment of the beam on the
left side of the section to exhibit a downward concavity, and that
corresponds to a negative bending moment, according to the sign
convention for bending moment.
Shear Diagram
386 | Statics
distributed forces and do not replace them with
the equivalent point load.
3. Lined up below the free body diagram, draw a
set of axes. The x-axis will represent the location
(lined up with the free body diagram above), and
the y-axis will represent the internal shear force.
4. Starting at zero at the right side of the plot, you
will move to the right, pay attention to forces in
the free body diagram above. As you move right in
your plot, keep steady except…
388 | Statics
and a upwards force on the left. A visual of these forces
can be seen in the diagram to the right.
Moment Diagram
390 | Statics
To read the plot, you
simply need to take the find the location of interest
from the free body diagram above, and read the
corresponding value on the y-axis from your plot.
Positive internal moments would cause the beam to bow
downwards (think a smile shape) negative internal
moments will cause the beam to bow upwards (think a
frown shape). You can also see the positive and negative
internal moments in the figure to the right.
Answer:
Support reactions.
392 | Statics
reactions at that support, namely By, Bx, and MB, as
shown in the free-body diagram in Figure 4.4b. Applying
the conditions of equilibrium suggests the following:
394 | Statics
bending moment is parabolic, the bending moment
diagram is a curve. In addition to the two principal
values of bending moment at x = 0 m and at x = 5 m, the
moments at other intermediate points should be
determined to correctly draw the bending moment
diagram. The bending moment diagram of the beam is
shown in Figure 4.5d.
Example 3
396 | Statics
Example 5
398 | Statics
Source: Internal Forces in Beams and Frames,
Libretexts. https://eng.libretexts.org/Bookshelves/
Civil_Engineering/
Book%3A_Structural_Analysis_(Udoeyo)/01%3A_Chap
ters/
1.04%3A_Internal_Forces_in_Beams_and_Frames
• +V means increasing M
• -V means decreasing M
• When V = 0, that’s max or min M
400 | Statics
How does each plot start/end? Reactions only if no applied load/
moment at ends:
• Cantilever:
402 | Statics
There are a few online programs that can help confirm the shape
that you found or help you learn how to translate loads into shear
and moment diagrams. These are not acceptable to use on the
exam or in homework and have limited free versions. This is not an
endorsement of any of the sites, just showing learning tools.
Key Takeaways
404 | Statics
6.3 Examples
Here are examples from Chapter 6 to help you understand these
concepts better. These were taken from the real world and supplied
by FSDE students in Summer 2021. If you’d like to submit your own
examples, please send them to the author [email protected].
1. Problem
Sketch:
2. Draw
406 | Statics
Free-body diagram:
Knowns
• w = 220 N/m
• OA = 0.5 m
• AC = 0.2 m
• AB = 0.4 m
• L = 1.9 m
4. Approach
5. Analysis
$$w=\frac{F}{L}\\F=wL\\F_R=220N/m\cdot
1.9m\\F_R=418N\\\sum F_X=0=B_X$$
$$\sum M_A=0=B_y(0.4m)-
F_R(0.55m)\\(0.4m)B_y=418N(0.55m)\\B_y=\frac{229.
9 N\cdot m}{0.4m}\\B_y=574.75N$$
$$\sum F_y=0=-F_R+A_y+B_y\\A_y=F_R-
B_y\\A_y=418N-574.74N\\A_y=-156.75N$$
Make a cut at C:
408 | Statics
0.475m)\\M_c=-156.75 N (0.2m)-(220N/m\cdot 0.95m\
cdot 0.475m)\\M_c=-130.625N\cdot m$$
6. Review
1. Problem
410 | Statics
2. Draw
Sketch:
Free-body diagram:
Knowns:
4. Approach
5. Analysis
(Ax, Cy)
\begin{aligned}
\sum F_{x}=0=A_{x}=0 \\
\sum M_{A}=0 &=-F_{1} \cdot 2 m-F_{2} \cdot 5.5
m+C_{y} \cdot 8 m \\
C_{y}=&+F_{1} \cdot 2 m+F_{2} \cdot 5.5 m \\
C_{y} &=\frac{500 N \cdot 2 m+300 N \cdot 5.5 m}{8
m} \\
C_{y} &=331.25 \mathrm{~N}
\end{aligned}
(Ay)
\begin{aligned}
\sum F_{y}=0 &=A_{y}+C_{y}-F_{1}-F_{2} \\
A_{y} &=F_{1}+F_{2}-C_{y} \\
A_{y} &=500 \mathrm{~N}+300 \mathrm{~N} – 331.25
\mathrm{~N} \\
A_{y} &=468.75 \mathrm{~N}
\end{aligned}
Cut 1: at B
412 | Statics
\begin{aligned}
\sum F_{X}=0=A_{X} &+N_{B}=0 \\
& N_{B}=0 \\
\sum F_{y}=0 &=A_{y}-V_{B}-F_{1} \\
V_{B} &=A_{y}-F_{1} \\
V_{B} &=468.75 N – 500 N \\
V_{B}=-31.25 N
\end{aligned}
\begin{aligned}
\sum M_{B}=& 0=-A_{y}(4 m)+F_{1}(2 m)+M_{B} \\
& M_{B}=A_{y}(4 m)-F_{1}(2 m) \\
& M_{B}=468.75 N(4 m)-500 N(2 m) \\
M_{B} &=875 \mathrm{~N} \cdot \mathrm{m}
\end{aligned}
\begin{aligned}
\sum M_{2}=0 =-A_{y}(5.5 \mathrm{~m})+F_{1}(3.5
\mathrm{~m})+M_{2} \\
M_{2} &=A_{y}(5.5 \mathrm{~m})-F_{1}(3.5
\mathrm{~m}) \\
M_{2} &=468.75 \mathrm{~N}(5.5 \mathrm{~m})-500
\mathrm{~N}(3.5 \mathrm{~m}) \\
M_{2} &=828.125 \mathrm{~N} \cdot \mathrm{m}
\end{aligned}
414 | Statics
Answer: NB = 0, VB = -31.25 N, MB = 875 Nm
6. Review
[latex]\bar{x}=\frac{m_1*x_1}{m_1+m_2}+\frac{m_2
*x_2}{m_1+m_2}[/latex]
420 | Statics
and that the sum of the weighting factors is always 1.
Also note that if, for instance, m1 is greater than m2,
then the position x1 of particle 1 will count more in the
sum, thus ensuring that the center of mass is found to
be closer to the more massive particle (as we know it
must be). Further note that if m1 = m2, each weighting
factor is 1/2, as is evident when we substitute m for
both m1 and m2 in equation 8-1:
$$\bar{x}=\frac{m}{m+m}x_1+\frac{m}{m+m}x_2\\\
bar{x}=\frac{1}{2}x_1+\frac{1}{2}x_2\\\bar{x}=\frac{x_
1+x_2}{2}$$
A second explanation:
422 | Statics
This idea is not limited just to two point masses. In
general, if n" role="presentation">𝑛 masses,
m1,m2,…,mn," role="presentation">𝑚1, 𝑚2,…,𝑚𝑛, are
placed on a number line at points x1,x2,…,xn,"
role="presentation">𝑥1,𝑥2,…,𝑥𝑛, respectively, then the
center of mass of the system is given by:
Example 1:
Solution
424 | Statics
In some sense, one can think about the center of mass
of a single object as its “average position.” Let’s consider
the simplest case of an “object” consisting of two tiny
particles separated along the x-axis, as seen below:
$$\bar
x_{cm}=(\frac{m_1}{m_1+m_2})x_1+(\frac{m_2}{m_1+
m_2})x_2=\frac{m_1x_1+m_2x_2}{M_{system}}$$
$$\vec{r}_{cm}=\bar x_{cm}\underline{\hat{i}}+\bar
y_{cm}\underline{\hat{ j}}+ \bar
z_{cm}\underline{\hat{k}}\\=\frac{[m_1 x_1+m_2
x_2+…]}{M}\underline{\hat{i}}+\frac{[m_1y_1+m_2y_2
+…]}{M}\underline{\hat{ j}}+\frac{[m_1 z_1+m_2
z_2+…]}{M}\underline{\hat{k}}\\=\frac{m_1[x_1\
underline{\hat{i}}+y_1\underline{\hat{ j}}+z_1\
underline{\hat{k}}]+m_2[x_2\underline{\hat{i}}+y_2\
underline{\hat{ j}}+z_2\
underline{\hat{k}}]+…}{M}\\=\frac{m_1\vec r_1+m_2\
vec r_2+…}{M}$$
Example 2:
426 | Statics
• m2 = 6 kg placed at (1, 1)m, and
• m3 = 4 kg placed at (2, -2)m.
Solution
First we calculate the total mass of the system:
$$ m = \sum_{i=1}^3 m_i = (2 + 6 + 4) kg = 12 kg $$
Then we have
428 | Statics
aluminum. Further imagine that the percentage of lead
in the rod varies smoothly from 0% at one end of the
rod to 100% at the other. The linear density of such a
rod would be a function of the position along the length
of the rod. A one-millimeter segment of the rod at one
position would have a different mass than that of a one-
millimeter segment of the rod at a different position.
Example 3:
430 | Statics
= µL with L being the length of the rod. The problem
with this is, that µ varies along the entire length of the
rod. What value would one use for µ ? One might be
tempted to evaluate the given µ at x = L and use that,
but that would be acting as if the linear density were
constant at µ = µ(L). It is not. In fact, in the case at hand,
µ(L) is the maximum linear density of the rod, it only has
that value at one point on the rod.
[latex]m=\frac{bL^3}{3}[/latex]
$$m=\frac{0.650\
frac{kg}{m^3}(0.890m)^3}{3}\\m=0.1527kg$$
[latex]\bar{x}=\frac{bL^4}{4m}[/latex]
Key Takeaways
432 | Statics
Looking Ahead: The next section will look at how to
calculate the center of mass for a complex object.
We use the centroid tables that are listed below to combine the
locations of the centers of mass for each shape. Approximations are
needed as real life objects are rarely perfectly square or circular, but
if they are symmetric, it makes it easier to approximate.
The locations of the center of mass (rcm) are as follows. The
source for the images are from Jacob Moore et al.
http://mechanicsmap.psu.edu/websites/centroidtables/
centroids2D/centroids2D.html
[latex]Area = bh[/latex]
Rectangle
[latex]r_{cm} =
\left[\frac{b}{2},
\frac{h}{2}\right][/latex]
[latex]Area =
\frac{1}{2}bh[/latex]
Right
Triangle
[latex]r_{cm} =
\left[\frac{b}{3},
\frac{h}{3}\right][/latex]
[latex]Area =
\frac{1}{2}bh[/latex]
Triangle
[latex]r_{cm} =
\left[\frac{b}{2},
\frac{h}{3}\right][/latex]
Circle
[latex]r_{cm} = \left[0,0\
right][/latex]
[latex]Area = \pi
(r_o^2-r_i^2)
\\r_o=\text{outer
radius}\\r_i=\text{inner
Circular
radius}[/latex]
Annulus
[latex]r_{cm} = \left[0,0\
right][/latex]
[latex]Area =
\frac{\pi}{2}r^2[/latex]
Semicircle
[latex]r_{cm} =
\left[0,\frac{4}{3\pi}r\
right][/latex]
436 | Statics
[latex]Area =
\frac{\pi}{4}r^2[/latex]
Quarter
Circle
[latex]r_{cm}=\left[\frac{4
3\pi}r,\frac{4}{3\pi}r\
right][/latex]
Ellipse
[latex]r_{cm} = \left[0,0\
right][/latex]
To find the cm, select the appropriate shape from the above table.
.
.
.
.
.
438 | Statics
Use the equation to solve for rcm.
1. Define an origin
2. Split the object up into recognizable shapes
3. Find the center of mass (cm) of each shape from the origin
4. Calculate the mass of each part: [latex]\rho =
\frac{m}{V}[/latex] (To find the centroid, this step can be
skipped and only the area or volume is used).
5. Use the weighted cm equations to find the x cm. Repeat for y
440 | Statics
For the shape shown at the top, we can break it down
into a rectangle (1), a right triangle (2), and a circular
hole (3). Each of these simple shapes is something we
have listed in the centroid table to the right.
442 | Statics
This generalized formula to find the centroid’s 𝑥-
location is simply Area 1 times [latex]\bar x_1[/latex]
plus Area 2 times [latex]\bar x_2[/latex], plus Area 3
times [latex]\bar x_3[/latex], adding up as many shapes
as you have in this fashion and then dividing by the
overall area of your combined shape. The equations are
the same for the 𝑦-location of the overall centroid,
except you will instead be using [latex]\bar y[/latex]-
values in your equations.
446 | Statics
websites/A2_moment_intergrals/method_of_composite_parts/
methodofcompositeparts.html with pdfs and video solutions.
Key Takeaways
450 | Statics
include moment integrals in one dimension, two
dimensions, and three dimensions, moment integrals of
force functions, of areas/volumes, or of mass
distributions, first order or second order moment
integrals, and rectangular or polar moment integrals.
452 | Statics
integrals versus polar moments integrals. This is a
difference in how we define the distance in our moment
integral. Let’s start with the distinction in 2D. If our
distance is measured from some axis (for example the x-
axis, or the y-axis) then it is a rectangular moment
integral. If on the other hand the distance is measured
from some point (such as the origin) then it is a polar
moment integral.
Key Takeaways
454 | Statics
moment of inertia), and an object’s resistance to rotating
(mass moment of inertia).
The red r’s in this image show the distance that is being measured
when adding up each little infinitesimal dm. Notice how the r
changees direction from x to y but looks the same between x and z.
Equations have been developed for common shapes so that you
don’t have to integrate every time you want to find the inertia of an
object. The result is different for each axis, as shown in the following
figure.
‘Ixx‘ can be read as ‘the inertia if rotating about the x-axis’. Notice
458 | Statics
particle, with respect to rotation about an axis from
which the particle is a distance r.
2
I = mr
I = I 1 + I2
2 2
I = m 1r1 + m 2r2
460 | Statics
7.4.2 Inertia Table of Common Shapes
$$ I_{xx} = \frac{1}{4}mr^2
\\I_{yy}=\frac{1}{2}mr^2 \\I_{zz} =
\frac{1}{4}mr^2 $$
Circular Plate
$$ I_{yy^\prime} = \frac{3}{2}mr^2
* thickness << 1
$$ I_{xx} = \frac{1}{12}m(3r^2+h^2)
\\I_{yy}=\frac{1}{2}mr^2 \\I_{zz} =
Cylinder \frac{1}{12}m(3r^2+h^2) $$
$$I_{xx}=\frac{2}{5}mr^2
\\I_{yy}=\frac{2}{5}mr^2
\\I_{zz}=\frac{2}{5}mr^2 $$
Sphere
462 | Statics
$$ I_{xx} = \frac{1}{12}ml^2
\\I_{yy}=0\\I_{zz} = \frac{1}{12}ml^2
Slender Rod
$$ I_{xx^\prime} = \frac{1}{3}ml^2
\\I_{zz^\prime} = \frac{1}{3}ml^2 $
* radius << length
$$ I_{xx} = \frac{1}{12}mh^2
\\I_{yy}=\frac{1}{12}m(h^2+b^2) \\I
Rectangular
\frac{1}{12}mb^2 $$
Plate
* thickness << 1
$$ I_{xx} = \frac{1}{12}m(h^2+d^2)
\\I_{yy}=\frac{1}{12}m(d^2+w^2) \\I
Rectangular = \frac{1}{12}m(h^2+w^2) $$
Block
$$ Volume = bwh $$
Asymmetric Shapes
$$I_{xx}=\frac{83}{320}mr^2
\\I_{yy}=\frac{2}{5}mr^2
\\I_{zz}=\frac{83}{320}mr^2 $$
Hemisphere
$$I_{xx^\prime}=\frac{2}{5}mr^2
\\I_{zz^\prime}=\frac{2}{5}mr^2 $$
$$I_{xx}=\frac{3}{80}m(4r^2+h^2)
\\I_{yy}=\frac{3}{10}mr^2
\\I_{zz}=\frac{3}{80}m(4r^2+h^2) $$
Cone $$I_{xx^\prime}=\frac{1}{20}m(3r^
2)
\\I_{zz^\prime}=\frac{1}{20}m(3r^2
$$
Hallow Shells
464 | Statics
$$ I_{xx} = \frac{1}{6}m(3r^2+h^2)
\\I_{yy}=mr^2 \\I_{zz} =
Cylindrical \frac{1}{6}m(3r^2+h^2) $$
Shell
* Thickness << 1
$$I_{xx}=\frac{2}{3}mr^2
\\I_{yy}=\frac{2}{3}mr^2
Spherical \\I_{zz}=\frac{2}{3}mr^2 $$
Shell
* Thickness << 1
$$I_{xx}=\frac{5}{12}mr^2
\\I_{yy}=\frac{2}{3}mr^2
\\I_{zz}=\frac{5}{12}mr^2 $$
Hemispherical
Shell
$$I_{xx^\prime}=\frac{2}{3}mr^2
\\I_{zz^\prime}=\frac{2}{3}mr^2 $$
* Thickness << 1
Notice how different objects with the same mass and radius
rotate at different rates. This simulation shows a cylinder (blue),
ring (green), solid sphere (yellow-brown), and spherical shell (red).
Which one has the least inertia? Why?
466 | Statics
[latex]\qquad I =
mk^2[/latex]
Source:
https://phys.libretexts.org/@go/page/
18431
For example, if the mass of an
object is m=10 kg, the radius of
gyration is 5 m, then the inertia is:
I = mk2 = 10 kg * 5 m * 5 m = 250 kgm2.
To find the radius of gyration:
$$ k=\sqrt{\frac{I}{m}}=\sqrt{\frac{250 kgm^2}{10kg}} = 5 m $$
Key Takeaways
468 | Statics
7.5 Inertia Intro: Parallel Axis
Theorem
There are two great uses for the parallel axis theorem:
1. Finding the inertia of a complex object with multiple parts.
To begin with, the parallel axis theorem is equal to the inertia about
the center of mass (Icm) plus the distance between the axes of
rotation squared times the mass.
$$I=I_{cm}+md^2$$
Example 1:
470 | Statics
Adapted from: Adapted from source: mechanicsmap.psu.edu/websites/
A2_moment_intergrals/parallel_axis_theorem/parallelaxistheorem.html
472 | Statics
shape.
7. Add up all these individuals inertias to find IT.
Example 2:
Example 3
Source: mechanicsmap.psu.edu/websites/A2_moment_intergrals/
parallel_axis_theorem/parallelaxistheorem.html
• For each sphere, the rcm is 0.3m + 1/2 radius = 0.3m + 1/2
(0.2m) = 0.4m
• For the bar, the rcm is at 0.
474 | Statics
Find the inertia about the center of mass for each shape separately.
Use the rectangle equation:
For the sphere:
$$ I_{cm-sph}=\frac{2}{5}mr^2 =\frac{2}{5}*(40kg)*(0.1m)^2 \\
\qquad \quad =0.16kgm^2$$
Finally, add up the parts: the 2 spheres and the cm of the rod:
$$I_{total} = 2 * I_{o-sph}+I_{cm-rod}\\\qquad \quad = 2*(6.56
kgm^2) + (0.6 kgm^2)\\\qquad \underline{I_{total} = 13.72
kgm^2}$$
Source: mechanicsmap.psu.edu/websites/
A2_moment_intergrals/parallel_axis_theorem/
parallelaxistheorem.html
Key Takeaways
476 | Statics
7.6 Examples
Here are examples from Chapter 7 to help you understand these
concepts better. These were taken from the real world and supplied
by FSDE students in Summer 2021. If you’d like to submit your own
examples, please send them to the author author [email protected].
1. Problem
Source:
https://encrypted-tbn0.gs
tatic.com/
images?q=tbn:ANd9GcTM
4e4xHaRSXBdQMGugm1g
ISi2Qgn7rQx_K3w&usqp
=CAU
478 | Statics
2. Draw
Known:
• mb = 0.45 kg
• db = 20 cm
• mL = 18 kg
480 | Statics
• mf = 8 kg
4. Approach
5. Analysis
$$I_{xx}=I_{yy}=I_{zz}=\frac{2}{3}mr^2\\\qquad
\quad=\frac{2}{3}(0.45
kg)(0.1m)^2\\I_b=0.003kgm^2$$
$$f=\frac{15cm}{2}\underline{\hat{i}}+\frac{7cm}{2}\
underline{\hat{ j}}+\frac{7cm}{2}\underline{\hat{k}}\\f
=(7.5\underline{\hat{i}}+3.5\underline{\hat{ j}}+3.5\
underline{\hat{k}})cm$$
482 | Statics
$$L_1=(3.5\underline{\hat{i}}+3.5\
underline{\hat{ j}}+15\
underline{\hat{k}})cm\\L_2=(15cm-7cm-1cm)\underlin
e{\hat{i}}+(0cm)\underline{\hat{ j}}+(7cm)\underline{\h
$$\\L=L_1+L_2\\L=(3.5cm+7cm)\underline{\hat{i}}+
(3.5cm+0cm)\underline{\hat{ j}}+(15cm+7cm)\underline{
\hat{k}}\\L=(10.5\underline{\hat{i}}+3.5\
underline{\hat{ j}}+22\underline{\hat{k}})cm$$
$$\underline{\underline{P}=(9.58\
underline{\hat{i}}+3.5\underline{\hat{ j}}+16.3\
underline{\hat{k}})cm}$$
484 | Statics
Part d (find the inertia of the person’s leg about point
A):
$$x_f=0.15m\\z_f=0.07m\\m_f=8kg$$
$$I_{ff}=\frac{1}{12}\cdot m\cdot
(x^2+z^2)\\I_{ff}=\frac{1}{12}(8kg)(0.15m^2+0.07m^2)
\\I_{ff}=0.0182kg\; m^2$$
$$
d^2=(0.0208m)^2+(0.128m)^2\\I_{pf}=I_{ff}+m_f(d_f)
^2\\I_{pf}=0.0182kgm^2+8kg[(0.0208m)^2+(0.128m)^
2]\\I_{pf}=0.1527kgm^2$$
$$r=0.035m\\h=0.3m\\m_L=18kg\\I_{LL}=\frac{1}{
12}\cdot m\cdot
(3_r^2+h^2)\\I_{LL}=\frac{1}{12}\cdot
18kg(3(0.035m)^2+(0.3m)^2)\\I_{LL}=0.1405kgm^2$$
$$r_{LP}=[(9.58-10.5)\underline{\hat{i}}+(3.5-3.5)\un
derline{\hat{ j}}+(16.3-22)\underline{\hat{k}}]cm\\r_{L
P}=(-0.92\underline{\hat{i}}-5.7\
underline{\hat{k}})cm\\I_{PL}=I_{LL}+m(d^2)\\I_{PL}
=0.1405kgm^2+18kg[(0.0092m)^2+(0.057m)^2]\\I_{PL}
=0.2005kgm^2$$
$$I_p=I_{PL}+I_{pf}\\I_p=0.1527kgm^2+0.2005kg\\
I_p=0.3532kgm^2$$
$$A=(10.5\underline{\hat{i}}+3.5\
underline{\hat{ j}}+37\underline{\hat{k}})cm\\P=(9.58\
underline{\hat{i}}+3.5\underline{\hat{ j}}+16.3\
underline{\hat{k}})cm\\r_{AP}=P-
A\\r_{AP}=[(9.58-10.5)\underline{\hat{i}}+(3.5-3.5)\und
erline{\hat{ j}}+(16.3-37)\underline{\hat{k}}]cm\\r_{AP}
=(-0.92\underline{\hat{i}}-20.7\
underline{\hat{k}})cm$$
$$I_A=I_p+m(d^2)\\I_A=0.3532kgm^2+(8kg+18kg)[(
486 | Statics
0.0092m)^2+(0.207m)^2]\\I_A=1.4695kgm^2\\\unde
rline{I_A=1.47kgm^2}$$
6. Review
1. Problem
488 | Statics
Annotat
ed from
original
source:
https://
commo
ns.wiki
media.o
rg/
wiki/
File:201
9_Inter
nationa
ux_de_
France
_Friday
_ladies
_SP_gr
oup_1_
Starr_A
NDREW
S_8D9A
6706.jpg
2. Draw
Knowns:
490 | Statics
• m = 60 kg
• ΣM = 200Nm
• h = 167cm
• d1 = 30cm
• d2 = 60cm
• t = 0.5sec
• Izz = ½ mr2
• ΣM = I∝
• ⍵0 = 0rad/sec
Unknowns:
• ∝1 = ?
• ∝2 = ?
• Izz1 = ?
• Izz2 = ?
• ⍵1 = ?
• ⍵2 = ?
4. Approach
5. Analysis
Part a:
Step 1: find the inertia when the arms are hugged to the
body
Izz1= 0.675 kg m2
Izz2 = 0.5(60kg)(0.3m)2
Izz2 = 2.7 kg m2
ΣM = I∝ therefore ΣM/I = ∝
∝1 = ΣM / Izz1
∝1 = 200 Nm / 0.675 kg m2
∝1 ≈ 296.296 rad/sec2
∝2 = ΣM / Izz2
∝2 = 200Nm/2.7 kg m2
∝2 ≈ 74.074 rad/sec2
492 | Statics
Part b:
⍵1 = 296.296 s-2(0.5s)
⍵1 = 148.148 rad/sec
⍵2 = 74.074 s-2(0.5s)
⍵2 = 37.037 rad/sec
6. Review
Physics Textbooks:
University Physics Volume 1:
https://courses.lumenlearning.com/suny-osuniversityphysics/
Introductory Physics : Building Models to Describe Our World
(pdf download): https://openlibrary.ecampusontario.ca/catalogue/
item/?id=4c3c2c75-0029-4c9e-967f-41f178bebbbb
Information at the foundation of modern science and technology
from the Physical Measurement Laboratory of NIST:
https://www.physics.nist.gov/cuu/Units/index.html
“UCD: Physics 9A – Classical Mechanics” by Tom Weideman,
LibreTexts is licensed under CC BY-SA.
Source: https://phys.libretexts.org/Courses/
University_of_California_Davis/
UCD%3A_Physics_9A__Classical_Mechanics
496 | Statics