Arce211structures 1 - +++

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 143

ARCE 211-Structures I

Lecture Notes
Winter Quarter 2006
Ansgar Neuenhofer
Assistant Professor, Department of Architectural Engineering
California Polytechnic State University, San Luis Obispo
Copyright 2006 Ansgar Neuenhofer (all photos by author except Fig. 3.22)

Structures

A reflection of culture, ingenuity and history

TABLE OF CONTENTS
1

Stabilizing objects in space-preventing motion


1.1 Introduction
1.2 Stabilizing a point in space
1.3 Stabilizing a line in space
1.4 Fixing a body in space
1.4.1 Stabilizing for gravity
1.4.2 Stabilizing for lateral motions

Statics of a point
2.1 Introduction
2.2 Point in 2-D
2.2.1 Resultant and components of 2-D forces
2.2.2 Rectangular components of 2-D forces
2.2.3 Equilibrium
2.3 Point in 3-D
2.3.1 Rectangular components of 3-D forces
2.3.2 Resultant of 3-D forces
2.3.3 Equilibrium
Problems

7
7
7
7
7
10
13
13
16
17
20

Statics of objects in two dimensions


3.1 Moment of a 2-D force
3.1.1 Introduction
3.1.2 Varignons theorem (Example 3.1)
3.2 Resultant of non-concurrent forces and moments
3.2.1 Introduction
3.2.2 Replace system of forces and moments by equivalent single force and single moment
(Example 3.2)
3.2.3 Replace the force-moment system by an equivalent single force
(Example 3.3)
3.2.4 Summary
3.3 Equilibrium
3.3.1 Reaction forces at supports
3.3.2 Free-body diagram
3.3.3 Equilibrium equations
3.3.4 Alternative equilibrium equations
3.3.5 Example 3.4
3.3.6 Example 3.5
3.3.7 Constraints, statical determinacy, statical indeterminacy
3.4 Structures with interior hinges
3.4.1 Discussion
3.4.2 Example 3.4
Problems

25
25
25
26
27
27

Statics of objects in three dimensions


4.1 Moment of a 3-D force
4.1.1 Discussion
4.1.2 Example 4.3
4.2 Resultant of parallel forces in 3-D
4.2.1 Discussion
4.2.2 Example 4.3
4.3 Equilibrium of parallel forces in 3-D
4.3.1 Discussion
4.3.2 Example 4.3
4.4 Equilibrium of general forces in 3-D
4.4.1 Discussion
4.4.2 Example 4.4
4.4.3 Example 4.5
Problems

47
47
47
49
50
50
50
51
51
51
53
53
53
55
56

1
1
1
2
3
4
4

28
30
30
31
31
32
33
34
35
36
37
38
38
39
40

Distributed loads, center of gravity, centroid


5.1 Introduction
5.2 Center of gravity, center of mass, centroid
5.2.1 Theory-principle of moments
5.2.2 Example 5.1
5.2.3 Example 5.2
5.3 Calculating centroids of composite areas
5.3.1 Discussion
5.3.2 Example 5.3
5.3.3 Example 5.4
5.4 Distributed load on beams
5.4.1 Discussion and introductory example (Example 5.5)
4.4.2 Example 5.6
Problems

Trusses
6.1 Introduction
6.1.1 Historical development
6.1.2 Load resisting behavior
6.2 Statical determinacy, indeterminacy and instability of trusses
6.3 Methods of joints
6.3.1 Procedure
6.3.2 Example 6.1
6.4 Special loading conditions
6.4.1 Discussion
6.4.2 Example 6.2
6.5 Methods of sections
6.5.1 Procedure
6.5.2 Example 6.3
6.6 Summary
6.7 Example 6.4
6.8 The 12-node structure
5.8.1 Building the model
5.8.2 Stabilizing the roof plane
5.8.3 Analyzing the roof plane (Example 6.5)
5.8.4 Trusses for vertical loads (remove interior columns)
Problems

79
79
79
79
85
86
86
87
90
90
91
91
91
92
93
93
95
95
97
100
106
107

7
7.1

117
117
117
118
121
121
121
122
122
123
123
123
125
127
127
128
134

7.2
7.3
7.4
7.5

Stress, strain, Hookes law and axially loaded structural members


Axial stress
7.1.1 Discussion
7.1.2 Examples
Axial strain
7.2.1 Discussion
7.2.2 Example 7.4
Relation between stress and strain
7.3.1 Tension Test
7.3.2 Hookes law, yield strength, ultimate tensile strength
Axial problems
7.4.1 Discussion
7.4.2 Examples
Statically indeterminate axial force problems
7.5.1 Discussion
7.5.2 Examples
Problems

60
60
61
61
62
64
66
66
67
68
69
69
71
72

1. Stabilizing objects in space-Preventing motion


1.1

Introduction

Forces acting on structures have the tendency to move them in certain directions (translations) and rotate them about
certain axes. Since we dont want our structure to move, we have to properly constrain the structures, thus preventing
the motions that the forces tend to impose. The process of preventing motion of an object or a structure by properly
fixing it to the ground or fixing it to other objects already at rest, is called stabilization. In this chapter, we use simple
pin-connected links to attach objects to the ground. A pin-connected link prevents relative displacements between its
two ends along the line of the link. The link constitutes a single reaction with either a positive (when it reacts in
tension) or a negative (when it reacts in compression) sense. When using these links we ignore the bending resistance
of the link. This is an important assumption.

1.2

Stabilizing a point in space

We start with stabilizing a point in space. A point in space has three possible motions (degrees-of-freedom), which are
the three translations in the x , y and z -directions. In Fig. 1.1, the first link is along the z -direction, which prevents
motion in that direction. The second link has a component in the x -direction. After building in the second link we have
prevented motion in the x -z plane, the only remaining motion is that along the y -direction. To prevent motion in the y direction, the third link must have a component along that direction.

link 3

z
y

possible motions
x
link 1

link 2

link 1

link 1
link 2

Fig. 1.1 Stabilizing a point in space using three links.


At this point it is important to mention that in all our discussion in this class, we are only concerned with the initial motion of objects. The exact or true motion that the point in space describes after installing a first link is spherical motion
about point A . Initially, however, when starting from a vertical position the z -component of that motion is small. We
can thus assume that the motion takes place in the x -y plane only. Fig. 1.2 illustrates this phenomenon by drawing a
projection in which the true motion is a circle and the initial motion tangential. The initial motion is always perpendicular to the link(s).

link

initial motion (tangential)


true motion (circular)
pin connection (glue)
provides no rotational resistance

Fig. 1.2 Initial and true motions.


From Fig. 1.1, we can develop the following important rule for preventing translation of an object:
We prevent a translation of an object along a given axis with a link that has a component along that axis.
1

1.3

Stabilizing a line in space

Unlike a point, whose motions are all translations, a line in space also has two rotational degrees-of-freedom in addition to the three translational motions. Since a line has infinitesimal dimensions in the two transverse directions ( y and z -directions), a rotation about the axis of the line (the x -axis in Fig. 1.3) does not change the position of the object
in space and thus is not a real motion.

Fig. 1.3 Possible motions of a line in space.


To prevent the five possible motions (three translations and two rotations) we need five links. By fixing one end of the
line, say A , with three links as a point in space (see Section 1.2) we have prevented all three translations such that the
two rotations about the y -and z -axes are the possible motions remaining (Fig. 1.4a). A fourth link in the vertical direction prevents the rotation about the y -axis (Fig. 1.4b). In order for the fifth link to prevent the remaining rotation
about the z -axis, it must have a component in the y -direction (see Fig. 1.4c).

3
1

(a)

1
2

(b)

3
1
2

(c)

Fig. 1.4 Stabilizing a line in space using five links.


Note that a fifth link without a component in the y -direction would not prevent rotation about the z -axis (see Fig. 1.5).
If the line rotates about the z - axis through point A , the initial motion of point B is in the y -direction such that the fifth
link is not engaged.

z
y

Fig. 1.5 Improperly constrained line in space (rotation about z -axis through point A not prevented).

From Fig. 1.4 and Fig. 1.5, we can develop the following important rule for preventing rotation of an object.
We prevent a rotation of an object about a given axis through a point with a link that is neither parallel nor concurrent with the given axis.
The vertical link 4 in Fig. 1.4 prevents rotation about the y -axis through point A since the link is neither parallel nor
concurrent with that axis. Link 4 does not prevent rotation about z -axis since it is parallel to that axis. Likewise, the
highlighted link in Fig. 1.5 about the z -axis since it is concurrent with the axis.

axis of rotation
z
A

Fig. 1.6 Improperly constrained line in space.


Since axes of instability dont always coincide with the coordinate axes we select, they are often difficult to detect. Fig.
1.6 shows an improperly constrained line since it has only four instead of the necessary five links. Since point A is
completely fixed, we know that the one motion we dont prevent is a rotation about an axis through point A . But what
is the direction of this axis? Since the highlighted link has a component along the y -axis the line cannot rotate about
the z -axis. Since the highlighted link has also a component along the z -axis the line cannot rotate about the y -axis.
The possible motion is thus a rotation about an axis through point A . That axis has both a component along the y -axis
and a component along the z -axis. Its direction is parallel to the direction of the link (see Fig. 1.6).

1.4

Fixing a body in space

Fig. 1.7 Possible motions of a body in space.

Bodies in space may experience six independent directions, three translations and three rotations (see Fig. 1.7).
1.4.1
Stabilizing for gravity
We start he stabilization process by arranging links that fix the object against motion due to its own weight, i.e. due to
gravity. Since gravity acts in the vertical direction, the motions we have to prevent are the translation in the z - direction and the rotations about the x - and y -axes. It is intuitive to start with vertical links. A first link that fixes point A to
the ground prevents translation along the z -axis. The remaining motions due to gravity are the two rotations about
the x and y -axes passing through point A (see Fig. 1.8).

Fig. 1.8 Possible motions due to gravity with a link in the z - direction.
If a second link is used as shown in Fig. 1.9, we prevent the rotation about the y -axis. A rotation about the x -axis is
still possible. After installing a third link in the z -direction, the structure is properly constraint against gravity action
and thus able to carry its own weight.

z
y

Fig. 1.9 Possible motion (rotation about the x - axis) due to gravity with two links in the z - direction.
1.4.2
Stabilizing for lateral motions
With stabilizing the structure for gravity load, we have prevented three out of the six motion, the displacement in the
z - direction and the rotations about the x - and y - axes. The structure is still unstable in the lateral directions with translations in the x - and y - directions as well as rotation about the z - axis. Fig. 1.10 depicts these three motions.

Fig. 1.10 Possible lateral motions.


As a fourth link we select a diagonal member with a component in the x - direction thus preventing the translation in the
x - direction.

z
y

Fig. 1.11 Possible lateral motions with a link in the x - direction.


After adding another diagonal with a component in the y - direction, y - translation is also prevented and the only
remaining motion is the rotation in the x -y plane, i.e. about the z -axis as shown in Fig. 1.12.

Fig. 1.12 Possible lateral motion with diagonal links in the x - and y - directions.

z
y

Fig. 1.13 Properly constraint structure with six links.


It is instructive to investigate the axis of instability if we now remove certain links from the properly constrained structure in Fig. 1.13. Say we look for the motion possible by removing link 2. Since point A of the plane is fixed in all three
translational directions by links 1, 4, 5, the resulting motion must be a rotation about an axis through point A . Moreover, link 3 prevents rotation about the x -axis. The axis of rotation is an axis through point A parallel to the direction of
link 6.

z
y

3
2
4

1
5

Fig. 1.14 Possible motion after removing link 2.

2. Statics of a point (particle)


2.1

Introduction

Forces acting on structures have the tendency to move them in certain directions (translations) and rotate them about certain
axes. Since we dont want our structure to move, we have to properly constrain the structures, that is we have to prevent the
motions the forces tend to impose. This chapter will introduce the properties and effects of forces and lays the foundation
for a fundamental understanding of statics. We should try to master this topic thoroughly.
A force can be defined as the action of one body on another and is a vector quantity since its effect depends on both the
magnitude and the direction and its point of application (a scalar quantity, in contrast, is defined by its magnitude only). In
SI units, the units of force are Newtons (N) and force is derived from the base units mass in kilograms (kg) and length in
meters (m), and time in seconds (s). One Newton is the force required to give a mass of 1 kg an acceleration of 1m/s2 .
m
1N = (1kg) 1 2
s

(2.1)

The weightW of a body is the force of gravity


W = mg

(2.2)

where m is the mass of the body and g=9.81 m/s2 is the gravity acceleration in SI units.
In US units, length in feet (ft), time in seconds (sec) and force in pounds (lb) are selected as base units. The mass of a body
is a derived quantity and is given by
m=

W
g

(2.3)

g=32.2 ft/sec2 is the acceleration of gravity in US units.


The units of force that we will use throughout this class are pound (# or lb), kilopound (kips or k=1000 #), Newton (N) and
kilonewton (kN=1000 N).

2.2
2.2.1

Point in 2-D
Resultant and components of 2-D forces

F1
R
F2
Fig. 2.1 Resultant and components.

We first look at the case of two concurrent forces, i.e. the forces act at the same point. Their direction is defined by the line
of action along with a sense given by an arrow. We can replace two forces F1 and F2 acting on a point by a resultant
force R that has the same effect on the point as the individual forces combined. In order to determine the resultant force, we
need to add vectors according to the parallelogram law. We cannot simply add the magnitudes, as we would do in a scalar
problem, since

R F1 + F2

(2.4)

unless the two forces are parallel. In that case, the magnitude of the resultant is the sum of the magnitude of the individual
forces F1 and F2 . Two forces F1 and F2 whose sum is the resultant R are called components of R .

2.2.2
Rectangular components of 2-D forces
Most commonly, we resolve a force into its rectangular components Fx and Fy . For the force in Fig. 2.2, the x - and y components are
Fx = F cos

Fy = F sin

(2.5)

They are related to the magnitude and direction of F by


= tan1

F = Fx2 + Fy2

Fy

(2.6)

Fx

Fy

Fx

Fig. 2.2 Rectangular components of force.


Resolving a force into its rectangular components is not always as straightforward as in Fig. 2.2. For example, angles need
not be measured counterclockwise from the horizontal axis. It is crucial that we are able to determine the correct components of a force regardless of how we orient the axes and measure the angles. In other words, we should not blindly rely
on such formulas as Fx = F cos and Fy = F sin .
When resolving forces into rectangular components, two things can go wrong:
(1) choose the wrong trigonometric function, (2) choose the wrong sign
To avoid the above mistakes, we should adopt the habit of determining the sign of the component first, then selecting the
correct angle between 0 and 90 degrees, and finally assigning the correct trigonometric function to the two components.
Note that by selecting an angle between 0 and 90 degrees, both the sin and cosine functions are positive. The following
examples should help us understand the above rules.
Problem: Find the components of the following forces of unit magnitude with respect to the given coordinate system
y
30

120D

210D

240D

D
D
D
Fx = F cos 60D = F sin 30D Fx = F cos 60D = F sin 30D
Fx = F cos 60D = F sin 30F
x = F cos 30 = F sin 60
D
D
D
D
D
D
D
D
Fy = F sin 60 = F cos 30Fy = F sin 30 = F cos 60 Fy = F sin 60 = F cos 30 Fy = F sin 60 = F cos 30

Fig. 2.3 Rectangular components of force.


In most cases, it is convenient to use rectangular components to find the resultant R of two or more forces. The x - component of the resultant R is the sum of the x - components of the individual forces. The y - component of the resultant R is
the sum of the y - components of the individual forces.
8

Rx = Fix

Ry = Fiy

i =1

(2.7)

i =1

Example 2.1
F1 = 4 N
30D
20D

F2 = 2 N

Fig. 2.4 Example 2.1.


Problem: Find the magnitude and direction of the resultant of F1 and F2 . Use (1) graphical, (2) trigonometric and (3) algebraic (using rectangular components) solutions.
Solution 1 (graphical solution)

F1 =4 N

R=5.5 N

14D

F2 =2 N

Fig. 2.5 Graphical solution.


We construct the parallelogram for vector addition of the two forces. Measurement of the length of R and direction yields
the approximate results
=14D

R = 5.5 N

(2.8)

Solution 2 (trigonometric solution)

20D

4
30

Fig. 2.6 Trigonometric solution.


We first calculate the angle
= 180D 30D 20D = 130D

(2.9)

and then apply the law of cosine which gives


R = 42 + 22 2 4 2 cos(130D ) = 5.503

(2.10)

To get the direction of R , we use the law of sines

4
5.503
4

=
= sin1
sin 130D = 33.83D
D

sin
sin 130
5.503

(2.11)

and finally
= 33.84D 20D = 13.84D

(2.12)

Solution 3 (algebraic solution):

F1

F1y

Ry

F2y

F2
F1x

F2x
Rx

Fig. 2.7 Algebraic solution.


(1) We calculate the x - and y - components of the forces F1 and F2 .
F1x

= 4 cos 30D = 3.464

F1y

= 4 sin 30D = 2.000

F2x

= 2 cos 20D = 1.879

F2y

= 2 sin 20D = 0.684

(2.13)

(2) We calculate the x - and y - components of the resultant force R by adding algebraically the components of the forces F1 and F2 .
Rx

= 3.464 + 1.879 = 5.343

Ry

= 2.000 0.684 = 1.316

(2.14)

(3) We find the magnitude of R by taking the square root of the sum-of-squares of the components (Pythagoras)
R = Rx2 + Ry2 = 5.3432 + 1.3162 = 5.503

(2.15)

(4) We find the orientation of R by


1.316
= 13.84
5.343

= tan1

(2.16)

Solution 3 is by far the most convenient and reliable method to find the resultant. From now on, we will use that
strategy exclusively.
2.2.3

Equilibrium

We now turn to the concept of equilibrium, the most central part of statics and key subject of this class. In this section, we
look at problems concerned with the equilibrium of points in two dimensions, i.e. equilibrium under concurrent forces in
2D. In the next section, we will extend the concept of equilibrium to bodies under action of non-concurrent forces.
Above, we learned how to determine the resultant of two or more concurrent forces. A special and important case arises,
when the resultant is zero. Then the point on which the forces act is in equilibrium and we define:
If the resultant of all forces acting on a point is zero, the point is in equilibrium.
The reverse is also true: If a point is in equilibrium, the resultant force acting on it must be zero.
In order for the resultant to be zero, both rectangular components of the resultant need to be zero. The conditions for equilibrium are thus
Rx =

=0

Ry =

=0

10

(2.17)

We explain and illustrate the strategy to solve an equilibrium problem by an example.


Example 2.2

30

40

C
F = 400 lb

Fig. 2.8 Example 2.2, Equilibrium problem.


Problem: Two cables are tied together as shown. Find the tension forces TAC and TBC in cables AC and BC , respectively.
Solution: Solving an equilibrium problem involves two fundamental steps, (1) the construction of a free-body diagram and
(2) the solution of the equilibrium equations.
(1) Free-body diagram
When solving an equilibrium problem, the first and most important step is to draw a free-body diagram. A free-body diagram shows the body (in this section, the body is a point) under consideration, i.e. the point at which the concurrent forces
intersect and all forces acting on the body (not by the body). Three forces act on point C , the unknown cable forcesTAC and TBC and the externally applied force F . Note that the forces exerted by the cables on point C are directed upward
(the cables are pulling on point C ). We should also include in the free-body diagram the geometric information necessary to
solve the problem (angles, distances, etc.). When drawing free-body diagrams throughout this set of notes, we use red color
for given (known) forces and blue color for unknown forces.
A
30

TAC

40

50

TBC
C

30
400

F = 400 lb

Fig. 2.9 Isolation of point C and corresponding free-body diagram.


(2) Equilibrium equations
The free-body diagram in Fig. 2.9 forms the basis for the equilibrium equations. We first sum up all the forces that act in
the x - direction and let their sum be zero. Since the x - components of the two cable forces go in different directions, they
have opposite sign. The externally applied force is vertical and thus has zero component in the x - direction so it does not
contribute to equilibrium in that direction. The equilibrium equation for the x - direction is thus
+ Fx = 0 = TAC cos 50 + TBC cos 30 = 0.643TAC + 0.866TBC

(2.18)

Next we sum up all force in the y - direction. Since the y - components of the two cable forces go in the same direction, they
have the same sign. The applied 400-lb force is directed downward and thus gets a negative sign.
11

+ Fy = 0 = TAC sin 50 + TBC sin 30 400 = 0.766TAC + 0.500TBC 400

(2.19)

Since both unknowns appear in both equations, we have to solve the equations simultaneously and obtain
TAC =352 lb

TBC =

(2.20)

261 lb

In case we dont remember how to solve simultaneous equations with two unknowns, we find a summary below. Let us
write the equations as
a11x 1 + a12x 2 = b1

(2.21)

a21x 1 + a22x 2 = b2

where x 1 and x 2 are the unknown variables, aij and bi are known coefficients. The two unknowns are
x1 =

b1a22 b2a12

x2 =

a11a22 a12a21

b2a11 b1a21

(2.22)

a11a22 a12a21

Substituting the numbers of our equilibrium problem yields the equations

0.643x 1 + 0.866x 2 = 0
0.766x 1 + 0.500x 2

(2.23)

= 400

and their solution

x1 =

0 0.5 400 0.866


= 351.7
0.643 0.5 0.766 0.866

x2 =

400 0.643 0 0.766


= 261.1
0.643 0.5 0.766 0.866

12

(2.24)

2.3

Point in 3-D

2.3.1

Rectangular components of 3-D forces

When working on problems involving forces in space we usually must find the x , y and z components of the forces. Often,
we define the direction of the force by either two points on the line of action of the force or by two angles, which define the
line of action.
(a) Specifying the line of action by two points
Consider a force acting at an arbitrary point A of a coordinate system with orthogonal axes x , y and z . We define the line of
action of the force, that is the direction of the force by a second point B . The objective is to resolve a force into its rectangular components Fx , Fy and Fz . We can visualize the relationship between a force and its three rectangular components by
drawing a box with edges parallel to the three coordinate axes. The diagonal AB of this box then represents the force F and
the components Fx , Fy and Fz of the force along the three coordinate axes are represented by the three edges.
B

z
y

z
L
Fz

Fx

Fy
y

Fig. 2.10 Components of a force in 3-D.


We observe that the ratios of the lengths of the three sides of the box to the length of the diagonal AB , say the length L , is
the same as the ratio of the components of the force to the magnitude of the force. In other words, the force components Fx , Fy and Fz and the distance components x , y and z are proportional.
F
x
= x
L
F

Fy
y
=
L
F

F
z
= z
L
F

(2.25)

If we know the coordinates of point A and B , the three orthogonal components of the force F are thus
Fx = F

x
=F
L

Fy = F

y
=F
L

Fz = F

z
=F
L

x
2

x + y 2 + z 2
y
2

x + y 2 + z 2
z
2

x + y 2 + z 2

x = x B x A
y = yB yA

(2.26)

z = z B z A

We note that by the Pythagorean theorem


F = Fx2 + Fy2 + Fz2

(2.27)

We may regard the expressions in Eq. (2.25) as the cosine of three angles (adjacent over hypotenuse) as shown in Fig. 2.11.

13

cos x =

F
x
= x
L
F

cos y =

Fy
y
=
L
F

cos z =

F
z
= z
L
F

(2.28)

z
y

x Fx
x

Fz

Fy
y

Fig. 2.11 Components of a force in 3-D


Since the three angles x , y , z are not independent, we can write an identity similar to Eq. (2.27)
cos2x + cos2y + cos2z = 1

(2.29)

(b) Specifying the line of action by two angles


We can also define a force in space by its magnitude and two angles. For example, 1 is the angle that F forms with the
projection of the force onto the x -y plane, and 2 is the angle that the projection forms with the x -axis.
z
y

F
z

1
Fz
y

Fig. 2.12 Resolving force F into components Fz and Fxy .

14

Fxy
2

Using 1 , we first resolve the force F into a horizontal component Fxy and vertical component Fz (see Fig. 2.12)
Fz = F sin 1

Fxy = F cos 1

(2.30)

z
y

z
1

Fz

Fxy
2

Fz

Fy

1
2

Fx
x

Fig. 2.13 Resolving force Fxy into components Fx and Fy .


We resolve the horizontal components Fxy further into x - and y - components (see Fig. 2.13)
Fx = Fxy cos 2 = F cos 1 cos 2

Fy = Fxy sin 2 = F cos 1 sin 2

(2.31)

The quantities Fx , Fy , and Fz are the rectangular components of F .


Example 2.3
z

3 ft

1ft

B
F

2 ft
y

3 ft

4 ft

Fig. 2.14 Example 2.3.


Problem: Point A is fixed by three links AB, AC , AD . Determine the components Fx , Fy , Fz of the force F acting on
support B and the angles 1 and 2 according to Fig. 2.12 and Fig. 2.13.
Solution: The force passes through A and B and is directed from A to B . The rectangular components of vector BA are
dx = 4 ft

dy = 3 ft

dz = 2 ft

(2.32)

and the distance from A to B is


AB = dx2 + dy2 + dz2 = 42 + 32 + 22 = 5.385 ft

(2.33)

Since the components of vector BA are proportional to the components of the force F , we can write

15

dx
4
F =
F = 0.7428 F (ans )
AB
5.385
dy
3
Fy =
F =
F = 0.5571 F (ans )
AB
5.385
d
2
Fz = z F =
F = 0.3714 F (ans )
AB
5.385
Fx =

(2.34)

It is always a good idea to check our math by using Eq. (2.27)


0.74282 + 0.55712 + 0.37142 = 1(ok)

(2.35)

The projection of the force F onto the x -y plane is


Fxy = Fx2 + Fy2 = 0.74282 + 0.55712 = 0.9285 F

(2.36)

such that angle 1 is


0.3714
= 21.80D (ans ) or 1 = cos1 (0.9285) = 21.80D (ans )
1 = tan1
0.9285

(2.37)

Angle 2 is
0.5571
0.7428
= 36.87 D (ans ) or 2 = cos1
= 36.87 D (ans )
2 = tan1

0.7428
0.9285

(2.38)

Fig. 2.15 illustrates the results by showing the original force and its components in three different projections.
x -y plane

0.5571F

0.9285F

0.5571F

x -z plane

B 0.7428F

0.3714F

0.3714F

0.7428F

y -z plane

0.6696F

x
0.8305 F

x
Fig. 2.15 Different projections of given force and its components.

Important:
When calculating components of a force in three dimensions, it is often advantageous to summarize the calculations in a
table:
x
y
z
L = x 2 + y 2 + z 2
x / L
y / L
z / L
4
3
2
5.385
0.7428
0.5571
0.3714
Note: In the table, we work with positive numbers only. We find the sign of the component by inspection.
2.3.2

Resultant of 3-D forces

Analogous to the procedure used in two dimensions, we calculate the components of the resultant of several concurrent
forces in space by summing up their orthogonal components
n

Rx = Fix
i =1

Ry = Fiy
i =1

Rz = Fiz

(2.39)

i =1

We obtain the magnitude of the resultant by the Pythagorean theorem


R = Rx2 + Ry2 + Rz2

(2.40)
16

The three angles x , y , z the resultant forms with the coordinate axes are
cos x =

Rx
R

cos y =

Ry

cos z =

Example 2.4

Rz
R

(2.41)

B
F2

F1

5m

3m
A

4m

Fig. 2.16 Example 2.4.

Problem: Two cables support a 10m tall post AB as shown. Knowing that the tension forces in the two cables
are F1 = 50 kN and F2 = 75 kN , find the three rectangular components of the resultant force R acting on the post at B .
Solution: In order to find the three rectangular components of the two cables forces, we set up the following table:
Cable
1
2

x
4
0

y
3
5

z
10
10

L = x 2 + y 2 + z 2
11.180
11.180

x / L
0.3578
0

y / L
0.2683
0.4472

z / L
0.8944
0.8944

Using Eq. (2.39), we obtain


Rx = Fx = 50 0.3578 + 0

= 17.89 kN (ans )

Ry = Fy = 50 0.2683 + 75 0.4472 = 13.415 + 33.540 = 20.31kN(ans )


Rz = Fz = 50 0.8944 75 0.8944 = 44.720 67.080 = 111.8 kN(ans )

2.3.3

Equilibrium

The three equations of equilibrium for concurrent forces in space (equilibrium of a point) are

=0

=0

=0

(2.42)

As for plane problems, the first step in solving a problem is always the construction of a free-body diagram. The
fundamental difference to 2-D problems is that we generally need to draw several free-body diagrams for different
projections of the 3-D problem.

17

Example 2.5
O
A
1

3
2

3m

z
x

B
1.5 m

2m
1.5 m

2m
C

Fig. 2.17 Example 2.5


Problem: A steel plate with mass 500kg is supported as shown. Find the tension force in the cables AB, AC , AD . Note that
the force in the vertical cable AO equals the force of gravity of the plate.
Solution:
To express the rectangular components of each force in terms of the unknown magnitudes F , we proceed analogously to
Examples 2.3 and 2.4. All three cables have the same length
AC = AB = AD = L = 22 + 1.52 + 32 = 3.905
x
1.5
=
= 0.3841
L
3.905

y
2
=
= 0.5121
L
3.905

0.38412 + 0.51212 + 0.76822 = 1.000

z
3
=
= 0.7682
L
3.905

(2.43)

(ok)

When can now express the three components of the force in each cable in terms of their unknown magnitudes F1, F2 , F3 by
F1x = 0.3841 F1

F1y = 0.5121 F1

F1z = 0.7682 F1

F2x = 0.3841 F2

F2y = 0.5121 F2

F2z = 0.7682 F2

F3x = 0.3841 F2

F3y = 0.5121 F3

F3z = 0.7682 F3

(2.44)

Analogous to the previous example, we can summarize the results in a table:


Cable
1
2
3

x
2
2
2

y
1.5
1.5
1.5

z
3
3
3

L = x 2 + y 2 + z 2
3.905
3.905
3.905

x / L
0.5121
0.5121
0.5121

y / L
0.3841
0.3841
0.3841

z / L
0.7682
0.7682
0.7682

The next step is to draw a free-body diagram. When analyzing three-dimensional problems, the free-body diagram serves
the same purpose as in two dimensions. We can either draw a three-dimensional view of the isolated body or show
projections along the coordinate axes. Whether we use one or the other or both is a matter of personal preference. Although
the sense of the unknown forces on the free-body diagram is arbitrary, we should try to represent the forces in their correct
direction whenever possible. For example, in this problem, all cable forces are tension forces and thus act away from
point A . We start with projecting point A and the forcing acting on it onto the x -z plane (we are looking along the
positive y -direction, Fig. 2.18). The weight of the plate is W = 9.81 m/s2 500 kg = 4905 N .

18

O
4905 N
A

4905 N

3
2

F1

x
D

B
1.5 m

F2

F1x A
F1z

3m

F2x

F3

F3x

F2z
F3z

2m
1.5 m

2m
C

Fig. 2.18 Free-body diagram (3-D view and projection onto x -z plane).
Using this free-body diagram, we are able to write two equations of equilibrium, one for the x -direction, the other for the z direction. The y -components of the three cable forces dont appear on the free-body diagram since they act perpendicular to
the viewing plane. Writing these two equations of equilibrium gives

F
F

= 0 = F1x F2x + F3x

= 0.3841F1 0.3841F2 + 0.3841F3

= 0 = Fz 1 F2z F3z + 4905

= 0.7682F1 0.7682F2 0.7682F3 + 4905

(2.45)

Next, we are looking at a free-body diagram obtained from a y -z projection (we are looking along the negative x -direction,
see Fig. 2.19). In this projection, the x -components of the three cable forces dont appear on the free-body diagram since
they act perpendicular to the viewing plane. This free-body diagram allows us to write an equation for equilibrium in the y direction.

4905 N
F3y

F2y A
F1z

z
F1y
y

F2z

F3z
Fig. 2.19 Free-body diagram (projection onto y -z plane).

= 0 = F1y F2y F3y

= 0.5121F1 0.5121F2 0.5121F3

(2.46)

Note that this free-body diagram would also allow us to write an equation of equilibrium for the z -direction, which is, of
course, identical to the equation above.
Summarizing the three equations of equilibrium, we obtain a set of three simultaneous equations

F
F
F

= 0 = F1x F2x + F3x

= 0.3841F1 0.3841F2 + 0.3841F3

= 0 = F1y F2y F3y

= 0.5121F1 0.5121F2 0.5121F3

= 0 = Fz 1 F2z F3z + 4905

= 0.7682F1 0.7682F2 0.7682F3 + 4905

Solving the equations, gives


F1 = 3192 N, F2 =0, F3 = 3192 N (ans )

19

(2.47)

Problems
2.1
y
B(2, 2)

F =10 kN
x

A(3, 1)

The line of action of the force F passes through points A and B . Find the x - and y - components of F .
Solution: Fx = 8.57 kN
2.2

Fy = 5.14 kN

F1
F2

At what angle must a force F2 = 400 lb be applied in order for the resultant R to have a magnitude of 1000 lb . The
force F1 has a magnitude of 700 lb .
Solution: = 51.32D
2.3
a
F
b

60D

50D

A force F has two components, one along a -a , the other along b -b . If the component along a -a has a magnitude
of 400 N , calculate the magnitude of the component along b -b . Also find the magnitude of F .
Solution: F = 452.2 N

Fb b = 490.7 N

20

2.4
y

2000 lb

F
800 lb

50

20D

One tension tie and two compression rods are attached to a support. Given the information that the resultant of the
three forces is downward with a magnitude of 3000 lb , calculate the magnitude of F and the angle .
=49.47 D

Solution: F = 2680 lb,


2.5

3m

2m

5m

2m

3m

The four stay cables of the cable-stayed bridge have forces F1 =250 kN, F2 =300 kN, F3 =350 kN, F4 =200 kN . Calculate the resultant R acting on the tower at A and its direction.
Solution:
R=

921.7 2 + 16.792 = 921.9 kN, = 88.9D (with respect to horizontal) (downward, to the left)
20 ft

2.6
50D
L1 =10 ft

W =500 lb

A weightW is suspended from two cables as shown. The length of cable 1 is L1 =10 ft . Find the tension forces in the
two cables.
Solution: T1 =442.9 lb, T2 = 326.9 lb

21

2.7

3 ft

F
36.87 D

C
A

2k
1 ft

Two cables are connected at A and loaded as shown. Find the range of values of F such that both cables remain taut.
2.8

100 k

55D

Determine the forces in links AB and BC as a function of the angle . Use EXCEL to plot the variation of the two forces for varying angle ( 0 90D ). Plot the angle on the x -axis and the forces on the y -axis and draw both forces
into a single plot.
2.9
A force F of magnitude 10 kips passes through points A and B with coordinates A(2, 3, 4) and B(5,1, 6) , respectively
and is directed from A to B . Find the x -, y -, and z components of F .
Solution: Fx = 7.276 k, Fy = 4.851 k, Fz = 4.851 k
z

2.10

y
B

F = 100 lb

10 in

A
30 in

20 in

O
A force F of magnitude 100 lb acts as shown. Find the x -, y -, and z components of F .
Solution: Fx = 80.18 lb, Fy = 53.54 lb, Fz = 26.73 lb

22

2.11
A force F of magnitude 100 kN has a component along the y - axis of Fy = 35 kN . The angle between the z - axis and
the line of action of F is = 30D . Find the x - , and z - components of F .
2.12
F4

F1
F2

F3

Four forces with the same magnitude F are concurrent at point A . The direction of the forces is defined by the box
above with dimensions of 4, 3, 2 along the x , y and z axes, respectively. Find (1) the components of the resultant R
along the x , y and z axes (in terms of F ), (2) the magnitude of R (in terms of F ).
Solution: (1) Rx = 2.437F , Ry = 1.989F , Rz = 1.373F
2.13

(2) R = 3.432F

3 ft

2 ft

B
A

2 ft

3 ft

4 ft

C
x
F

Two cables AC and BC support a beam CD . If the tension force in cable BC is FBC = 500 lb and the resultant of the
three forces FAC , FBC and F is directed along CD , find the tension force in cable AC . Also find the force F .
Solution: FAC = 750 lb, F = 603.5 lb

23

2.14

9 ft
1

6 ft

B
6 ft

A force F = 10 k is applied to point D of the above tripod. Find the forces in the three legs, if F is a force (a) in the x direction, (b) in the y -direction, (c) in the negative z -direction, (d) in the x -y plane with equal components in the x and y -directions, (e) in the x -z plane with equal components in the x - and negative z -directions, (f) in the y -z plane
with equal components in the y - and negative z -directions, (g) with equal components in the x , y - and negative z -directions. Clearly indicate whether the legs are in tension or compression.
Solution:
Link
(a)
(b)
(c)
(d)
1
15.00
15.00
10.000
0
2
18.03
0
0
12.75
3
0
18.03
0
12.75
Note: positive = tension, negative = compression, forces in kips
F

2.15

(e)
17.68
12.75
0

(f)

(g)

3.54
0
12.75

5.77
10.41
10.41

A
B

O
z
y

Repeat Problem 3.10 for the tripod above (height 9 ft ). Points A, B,C form an equilateral triangle in the x -y plane. The
projection O of point D onto the x -y plane is the center of a circle of radius 6 ft through A, B,C . The direction OA is
along the x -axis. Clearly indicate whether the links are in tension or compression.
Solution:
Link
(a)
(b)
(c)
(d)
(e)
(f)
1
12.02
0
4.01
8.50 11.33
2.83
2
6.01 10.41
4.01
3.11
1.42 10.19
3
6.01
10.41
4.01
11.61
1.42
4.53
Note: positive = tension, negative = compression, forces in kips

24

(g)
9.25
4.85
7.17

3. Statics of objects in two dimensions


3.1
3.1.1

Moment of a 2-D force


Introduction

In Chapter 2, we assumed that all forces were concurrent at a single point, a particle and the tendency of a force has been to
move the particle in a certain direction. Now we look at the effect of forces on bodies, i.e. we have to take the size of body
into consideration. Since forces acting on a body have different points of application, in addition to the tendency to move
the body, a force may also tend to rotate a body about an axis. The tendency of a force to rotate is called the moment of a
force.

(a)

(b)

d
a

F
(c)

(d)

A
a

A
a

d
a

Fig. 3.1 Moment of a force.


To motivate the idea of a moment we compare the two column-footing designs below. In Fig. 3.1a and Fig. 3.1c, the
structure is symmetrical, i.e. the column and footing centerlines are the same. In Fig. 3.1b and Fig. 3.1d, the column and
footing centerlines are different, i.e. the column has an eccentricity with respect to the footing.
In Fig. 3.1c and Fig. 3.1d we have replaced the two columns by corresponding forces acting on the foundations. In case (c),
the force has the tendency to move the foundation downward only. In case (d), the force has the tendency to move the foundation downward and rotate the foundation about an axis perpendicular to the plane of the paper. We measure the tendency
to make a body rotate about an axis through a point by the moment of the force about that axis. For example, the magnitude
of the moment M A of the force F about point A is defined as the product of the magnitude F of the force and the perpendicular distance d from A to the line of action of F .
MA = F d

(3.1)

25

The perpendicular distance is often called moment arm. If we use SI units, the moment has units of Newtonmeters or kilonewtonmeters (Nm, kNm). In the US system, the moment has units (lb-ft, lb-in, k-ft, k-in). A moment has either a clockwise
or a counterclockwise orientation depending on the relative position of the force and the axis. In two-dimensional problems, we consider a counterclockwise moment as positive and a clockwise moment as negative (see Fig. 3.2). It is
essential to work sign-consistent within a given problem.
M A positive
A

M A negative

M A positive

M A negative

Fig. 3.2 Sign of moment.


In this class, we restrict our discussion to two-dimensional problems, i.e. we deal with forces that act in a given plane, say
the x -y plane. When dealing with coplanar forces, the specification of the moment axis is redundant since all moments are
about the z - axis. Thus we commonly speak of the moment about a point. What we are actually implying is a moment about
an axis normal to the plane (the z - axis) passing through the point.
3.1.2

Varignons theorem

The French mathematician Varignon (1654-1722) developed an important theorem of statics, which states:
A moment of a force about a point is equal to the sum of the moments of its components about the same point.

Fx
Fy
2.89

F
20D
A

30D

30D
5.00

5.00

Fig. 3.3 Moment M A using Varignons principle.


To illustrate Varignons principle by an example (Example 3.1), we calculate the moment of the force F about point A using
two approaches:
(1) Varignons theorem
(2) the original definition of the moment as force times moment arm.
(1) The rectangular components of the force are (see Fig. 3.3)
Fx = F sin 20D = 0.3420 F

Fy = F cos 20D = 0.9397 F

(3.2)

The moment arm dy for the horizontal component Fx is the perpendicular distance from point A to Fx , hence
dy = 5.00 tan 30D = 2.887

The moment arm dx for the vertical component Fy is the perpendicular distance from point A to Fy , hence
26

(3.3)

dx = 5.00

(3.4)

such that the moment about A produced by the components of F becomes


M A = Fx dy Fy dx = 0.3420 F 2.887 0.9397 F 5 = 0.9874F 4.6985F = 3.711 F

(3.5)

We should note that in applying Varignons principle, we must properly account for the sign (clockwise or counterclockwise) of each moment. As defined earlier, we consider a counterclockwise moment positive and a clockwise moment
negative. The horizontal component Fx has the tendency to rotate counterclockwise about A . The moment generated by Fx is
thus positive. The vertical component Fy has the tendency to rotate clockwise about A . The moment generated by Fy is thus
negative.
(2) For the original definition M =Fd of the moment, we need to find the perpendicular distance d from the reference
point A to the line of action of F . Applying the rules of trigonometry, we have (see Fig. 3.4)
d=

5
cos 50D = 3.711
cos 30D

(3.6)

such that the moment produced by the force F is


M A = F 3.711

(3.7)

which coincides with the result of (1). Note that the force F has the tendency to rotate clockwise about A such that the sense
of M A is negative.
5.00

5 / cos 30D

F
20D

30D

50D

A
d

90D

Fig. 3.4 Moment M A using original definition.


In most cases, Varignons principle is an easier procedure to find the moment of a force about a point than the original definition. This is because finding the perpendicular distance sometimes requires involved geometric calculations.

3.2

Resultant of non-concurrent forces and moments

3.2.1
Introduction
In Chapter 2, we learned how to calculate the resultant force by adding all forces acting on a particle. In this Chapter, we
look at forces and moments acting on two-dimensional bodies and we will learn how to describe their resultant. The resultant effect is usually not only a resultant force, as was the case for particles, but a resultant force and a resultant moment.
We will perform two basic analytical procedures. (1) We will replace a given system of forces and moments by an equivalent system of a single force and a single moment at a given point. (2) We will calculate the point at which we can replace
a given system of forces and moments by a single force only.

27

(a)

A
F

(b)

(d)

M =Fd

(c)

Fig. 3.5 Replacing a force by a force at a different location and a moment.


Fig. 3.5 illustrates the resultant effect with respect to point A of a force acting at B . We can replace a force acting at
point B by the same force at point A and the moment M =Fd . Note that in Fig. 3.5(b) we add equal and opposite forces F without changing the net effects on the body. What we learn from Fig. 3.5 is that we can move a force acting on a rigid body from one point to another provided we add the correct moment. That moment compensates for the change in the
line of action of the force. Furthermore, we observe that we can view a moment as two equal and opposite forces F with parallel lines of actions that are distance d apart (see Fig. 3.6).
F =M / d

F =M / d
Fig. 3.6 Illustration of moment.

3.2.2

Replace system of forces and moments by equivalent single force and single moment (Example 3.2)
100 N

4.00

30D

45D

200 N

80 Nm

5.00

[m]

5.00

Fig. 3.7 Force-moment system (Example 3.2).


Problem: The rigid body in Fig. 3.7 is subjected to two forces and a moment. Replace the given force-moment system by a
single force and a moment at point A .
Solution:
(1) We will move all the forces to the single point A, and add the necessary moment such that the end product will be a
single force and a single moment at A. The magnitude of the moment is equal to the sum of the moments of the individual
forces about point A. Note we will have to add the externally applied moment of magnitude 80Nm. Hence the reduced system of forces consists of a force passing through A having the components
( +) Rx

= 200 cos 45 + 100 cos 30 = 54.82 N

( +) Ry

= 200 sin 45 + 100 sin 30

(3.8)

= 191.4 N

28

Instead of defining the force by its two rectangular components describe it by the magnitude
R= 54.82 2 + 191.4 2 = 199.1 N

(3.9)

and the angle with the x-axis


191.4
= 74.02D
= tan1
54.82

(3.10)

The corresponding moment (positive in the counterclockwise direction) is

M A = 80 + 200 sin 45 5.00 100 cos 30 4.00 + 100 sin 30 10.00 = 780.7 Nm

4.00

(3.11)

199 N

191N
54.8 N
74DA

781Nm
5.00

5.00

Fig. 3.8 Equivalent single force-single-moment system.


Remarks:
We have used Varignons theorem to compute the moment of the forces as the sum of the moments of their components.
The externally applied moment is included in this calculation. The point of application of the moment need not be
given, as the moment is a free vector.
Since the result for Rx is negative, the orientation is opposite to that assumed as positive in Eq. 3.8.
We should try to concentrate hard when deciding on the sense of the moments about A produced by the
components of the applied forces. The vertical components of both forces tend to rotate counterclockwise
(positive sign), whereas the horizontal component of the 100-N force tends to rotate clockwise (negative sign).
The horizontal component of the 200-N force passes through reference point A and thus has zero moment about
that point.
We can determine the resultant force either by its two rectangular components (red arrows in Fig. 3.8) or by its
magnitude and direction (blue arrow in Fig. 3.8).

29

Replace the force-moment system by an equivalent single force (Example 3.3)

4.00

3.2.3

199 N
191 N

54.8 N
A

x 0 = 4.09

74D

5.00

5.00

Fig. 3.9 Equivalent single force system (Example 3.3).


We now wish to replace the force-moment system calculated above with a single force. The distance x 0 should be such that
the moment of R about point A is equal to 781Nm in the counterclockwise direction. The easiest way to find the moment
of R is to resolve it into components at the intercept point. The x - component has zero moment about A, and hence
x0 =

781 Nm
M
=
= 4.09 m
191 N
Ry

(3.12)

Hence, the system of forces has been replaced by or reduced to a single resultant as described above. Note that we ignore
any sign in Eq. 3.12 and let the physics determine the placement of R . Had M A been clockwise, we would have had to
move R in the other direction.
Note that as an alternative we could have defined the line of action by the intercept with the y - axis.
3.2.4

Summary

(1) We can replace, at any location, an arbitrary number of forces and moments applied to a rigid body by a single force and
a moment. Note that the magnitude of the force does not depend on the location but the magnitude of the moment does.
(2) We can replace, at a specific location, an arbitrary number of forces and moments applied to a rigid body by a single
force.
Why are the three systems in Figures Fig. 3.7, Fig. 3.8 and Fig. 3.9 equivalent?
The systems are equivalent because the sums of the x - components, of the y - components, and of the moments about any
point are, respectively, equal.
or
The systems are equivalent because they tend to impart to the given rigid body the same translation in the x - direction, the
same translation in the y - direction and the same rotation about any point.
We will return to the principle applied in Sections 3.2.2 and 3.2.3 in Chapter 5 when calculating the center of gravity of an
object.

30

3.3

Equilibrium

A rigid body is in equilibrium if and only if the external forces and moments acting on it form a system of forces equal to
zero, i.e. a system that has zero resultant force and zero resultant moment. Differently put, a body is in equilibrium if the external forces do not impart any translational motion or rotation on the body. Hence, we have the equilibrium equations

=0

=0

=0

(3.13)

For the moment equation, we can select as the reference point A an arbitrary point, which does not have to be on the body.
When forces act on a structure (the action forces), we need to have a set of reaction forces such that both sets of forces are
in equilibrium, i.e. the resultant force and resultant moment of the applied forces and the reaction forces must be zero. Differently put, reaction forces are forces that develop to resist the tendency of the applied forces to translate and rotate a body.
An equilibrium problem that we are concerned with in this class is thus typically posed as follows:
Given a set of externally applied forces (and moments), find the reaction forces (reaction moments).
Structures can be supported in many different ways some of which we will briefly discuss in the following section.

Picture 3.1 Moment equilibrium-Large force times small moment arm = small force times large moment arm.
3.3.1

Reaction forces at supports

The support conditions that we deal with in this class, fall into three categories (see Fig. 3.10 and Picture 3.2):
(1) Reaction with a force with known line of action. We commonly refer to this support condition as a roller. This support
prevents motion in one direction only and involves one unknown.
(2) Reaction with a force of unknown direction. We typically call this support a pin support or pin connection. A pin support prevents translation of the body in both directions, but cannot prevent the body from rotating about the connection. This
support involves two unknowns usually represented by the x - and y - components of the force.
(3) Reaction with a force of unknown direction and a moment typically called fixed, clamped or built-in support. This
support prevents motion completely, i.e. two motions and rotation and involves three unknowns, the two components of the
force and the moment.

31

Picture 3.2 shows roller and pin supports of a steel bridge and a typical moment connection
(a)

(b)

(c)

Picture 3.2 (a) roller, (b) pin, (c) fixed supports.


We use the symbols in Fig. 3.10 to describe the support condition in our problem sketches

Fig. 3.10 Symbols used for support conditions. (a) roller, (b) pin, (c) fixed supports.
3.3.2

Free-body diagram

As for equilibrium of points discussed before, it is essential to draw a correct free-body diagram of the body under
consideration. We recall from Chapter 2 that a free-body diagram clearly defines the body under consideration and all forces acting on that body and is the physical model to which we apply the equilibrium equations.
Fist we have to make a decision regarding the free body that we use. We then separate the body from the ground or from
other bodies and indicate clearly all forces acting on the body. These forces are usually externally applied forces or gravity
forces that we know and reaction or support forces that we dont know. As mentioned earlier, we use blue color for unknown forces in the free-body diagrams, i.e. forces to be determined solving the equilibrium equations, and red color for
known forces, i.e. forces whose magnitude is known to us. A correct free-body diagram paves the way for a successful and
straightforward application of the equilibrium equations.

32

B
F

90D
mass m
frictionless contact at B

F
W =mg

Ax

Ay

Fig. 3.11 Mechanical system and free-body diagram.


Fig. 3.11 shows an example of a mechanical model along with the corresponding free-body diagram. We designate the
weight of the beam, which acts through its center of mass (assumed known). The support at A is a pin support, which is represented on the free-body diagram by forces in the horizontal and vertical directions with the orientation of the components
assigned arbitrarily. The frictionless contact at B exerts a normal force on the beam, i.e. the line of action of the force is perpendicular to the beam.
CONSTRUCTION OF THE FREE-BODY DIAGRAM
IS
THE MOST IMPORTANT STEP IN THE SOLUTION OF STATICS PROBLEMS

3.3.3
Equilibrium equations
In two-dimensional problems, we apply Eq. (3.13) to four categories of equilibrium, which are illustrated in Fig. 3.12.

Collinear

=0

= 0, Fy = 0

= 0, M z = 0

= 0, Fy = 0, M z = 0

Concurrent at a point

x
y
Parallel

x
y
General

Fig. 3.12 Equilibrium Categories.

33

3.3.4

Alternative equilibrium equations

It is not necessary to use the equilibrium equations in the form of Eq. (3.13). There are two alternative ways of requiring
equilibrium in two dimensions:
(1) If we require M P =0 about an arbitrary reference point P , the resultant force must be a force passing through P . If,
in addition, MQ =0 holds about a second reference pointQ , the resultant force must be a force passing through Q . That
force, however, is zero and equilibrium is satisfied if we also require M R =0 , where reference point R must not lie on
the line PQ . We thus have an equivalent equilibrium condition

=0

=0

=0

(3.14)

where the three reference points must not lie on a straight line. Fig. 3.13 illustrates this equilibrium condition.

F Q
P

P
F =0

=0

=0, MQ =0

Q
R

=0, MQ =0, M R =0

Fig. 3.13 Three moment equations for equilibrium.


(2) If we require M P =0 about an arbitrary reference point P , the resultant force must be a force passing through P . If,
in addition, Fx =0 holds along any arbitrary direction x , a potential force F has no other choice than acting through P
and perpendicular to x . If we select a second reference pointQ , the resultant force must vanish if the direction PQ is not
perpendicular to x . We thus have an equivalent equilibrium condition

=0

=0

=0

(3.15)

where PQ must not be perpendicular to the x - direction. Fig. 3.14 illustrates this equilibrium condition.
x

90D

=0

90D
F =0

=0, Fx =0

Fig. 3.14 One force equation and two moment equations for equilibrium.

34

x
M
=
0,
F
=
0,
p x MQ =0

3.3.5

Example 3.4

100 kN
20 kN

6.00

6.00

100 kN
20 kN

[m]
A

2.50

Ax

Ay

2.50

By
2.50

2.50

Fig. 3.15 Example 3.4.


Problem: Two forces are applied to the structure in Fig. 3.15, which is supported by a pin at A and a roller at B . Find the
reaction forces at supports A and B .
Solution:
(1) First, we draw a free-body diagram of the structure. The support at A is a pin connection that can resist a force in any
direction. We thus represent the reaction force at A by its components Ax and Ay . The support at B is a roller connection; the
support force is vertical and denoted by By . Thus a total of five forces act on the structure, which have to be in equilibrium.
(2) Next, we formulate equilibrium. When writing equilibrium equations we should always try to choose equations containing only one unknown to avoid solving simultaneous equations.
(1)

( +) Fx

= 0 = 20 + Ax

Ax = 20.0 kN

(2)

(cc+) M A

= 0 = By 5 100 2.50 20 6.00

By = 74.0 kN

(3)

( +) Fy

= 0 = Ay + By 100 = Ay + 74 100

Ay = 26.0 kN

(3.16)

We can check the results by taking moments about B

= 0 = 26 5 100 2.50 + 20 6.00

ok

(3.17)

Remarks:
We have arbitrarily assigned a direction to forces Ax , Ay and By .
The direction selected as positive in the equations of equilibrium (Eq. 3.16) is also arbitrary and independent of the
direction of the forces selected in the free-body diagram. An equation
( +) Fx

= 0 = 20 Ax

Ax = 20.0 kN

(3.18)

expressing horizontal equilibrium would be equally correct.


The negative result for Ax is indicating that the sense of that reaction force is opposite to that assumed in the freebody diagram.

35

Example 3.5

B
A

B
MA

1.00

[m]

30D

Ax

10 kN
[m]
5.00

2 kNm

30D
A
Ay

1.00

3.3.6

10 kN
5.00

2 kNm

Fig. 3.16 Example 3.5.


Problem: A force and a moment are applied at the free end of the cantilever structure in Fig. 3.16. Find the support reactions
at the fixed support A .
Solution:
First, we draw a free-body diagram of the structure. The support at A is a fixed connection that can resist a force in any direction and a moment. We thus represent the reaction force at A by its components Ax and Ay , and the reaction moment
by M A .
Next, we write the equilibrium equations.
(1)

( +) Fx

= 0 = 10 cos 30D + Ax

Ax = 8.66 kN

(2)

( +) Fy

= 0 = Ay 10 sin 30D

Ay = 5.00 kN

(3)

(cc+) M A = 0 = 10 sin 30 5.00 10 cos 30 1.00 + 2 + M A


D

(3.19)

M A = 31.66 kNm

Remarks:
We have used Varignons theorem to calculate the moment of the applied force about A . Both the x - and the
y - components of the 10-kN force tend to rotate about A in the clockwise direction. Consequently, the signs are
both negative.
The negative result for Ax is indicating that the sense of that reaction force is opposite to that assumed in the freebody diagram.
We have included the externally applied 2-kNm moment in the moment equation (3).

31.66 kNm

30D

8.66 kN

1.00

We can check the results by taking moments about B . It is good practice to redraw the free-body diagram and designate all
force by their numerical value. Since the numerical result for Ax is negative (see Eq. 3.19), we change the sense of the arrow
representing this force (see Fig. 3.17).

10 kN

5 kN

5.00

2 kNm

Fig. 3.17 Example 3.5, Equilibrium check.

= 0 = 2 5 5.00 + 31.66 8.66 1.00

ok

36

(3.20)

3.3.7

Constraints, statical determinacy, statical indeterminacy

In both examples discussed above we properly constrained the rigid body under consideration against any movement under
the given loads or any other load condition. Furthermore, we used no more constraints than necessary to prevent the
structures from moving. We call structures that have the minimum number of constraints necessary for equilibrium statically determinate. If we support a structure by more constraints than necessary for equilibrium, we say the structure is
statically indeterminate. The constraints that we can remove without destroying equilibrium are redundant constraints
and the number of redundant constraints represents the degree of statical indeterminacy. A structure whose supports are
arranged in a way that makes static equilibrium impossible is called improperly constrained or incompletely fixed. We
should avoid using improperly constrained structures in practice.
In this class, we only analyze statically determinate structures. However, it is important that we learn to correctly identify
statically determinate, statically indeterminate and improperly constrained structures. Fig. 3.18 presents examples of
statically determinate, statically indeterminate and improperly constrained structures.

statically determinate (simply supported beam)

statically determinate (cantilever)

improperly constrained (unstable)

improperly constrained (unstable)

statically indeterminate to the 1st degree

statically indeterminate to the 1st degree

Fig. 3.18 Different support conditions.


Structures 1 and 2 are both statically determinate since they have the minimum number of constraints necessary to prevent
all three motions of a two-dimensional body. Structure 3 has more than enough constraints to prevent vertical translation
and rotation but does not offer any resistance to horizontal movement, i.e. a horizontal force applied to Structure 3 cannot be
supported. Structure 3 is thus improperly, partially or incompletely constrained. Structure 4 provides constraints with
respect to vertical and horizontal motions and but offers no initial resistance to rotation, i.e. the beam can rotate about the
left support by a small angle. Hence, Structure 4 is also improperly constrained. Cases 5 and 6 are examples for statically
indeterminate structures. There are more constraints than necessary for equilibrium. For example, we could remove the
roller supports at the right end of the beam for Structures 5 and 6 without sacrificing the stability of the system. The same
holds for the center roller support in Structure 5.

37

3.4

Structures with interior hinges (pins)

3.4.1

Discussion

Designers often use interior/internal hinges (pins) to connect structural members. Consider the structure in Fig. 3.19. Segment AB of the structure is hinged to segment BD at the internal hinge B . A free-body diagram of the whole structure
shows four unknown reaction forces (see Fig. 3.20). The three equations of equilibrium that we can apply to the structure as
a whole are thus not sufficient for finding the reaction forces. We need additional equations to find the reactions.
F1

F2

Fig. 3.19 Beam with internal pin.

F1

F2

F2

F1

D
A

Dx
Dy

Fig. 3.20 Free-body diagram for whole structure.


To find the reaction forces we separate the structure at pin B and apply the two force components transmitted at B to
each part resulting in the two separate free-body diagrams shown in Fig. 3.21. It is important to note that there is no moment
transferred at connection B since a hinge by definition cannot transfer a moment from one part of the structure to another.
Also note that the forces at B obey the principle of action and reaction that is the sense of the two forces at B applied to
segment AB is opposite to that of the forces at B applied to segment BD . By separating the structure at the internal pin we
have increased the number of unknown forces from four to six. Applying the three equations of equilibrium to parts AB
and BD separately, however, results in six equations for calculating the six unknown reaction forces (see Fig. 3.21).

F1

FBD I
A

F2

Bx

Bx

By

By

FBD II
C

Dx
Dy

Fig. 3.21 Free-body diagrams of two beam segments.


We should note that the connections at B and D are statically identical. Pin B transmits a force of arbitrary direction (represented by its two components) from one part of the structure to another. Pin D transmits a force of arbitrary direction (again
represented by its two components) from the structure to the ground.
We also find interior hinges in arch structures, since designing an arch as two halves often simplifies construction. With two
hinges at the supports and one internal hinge, these structures are called three-hinged or three-pinned arches (see Fig. 3.22

38

Fig. 3.22 Three-pinned arches: structural system (left), High Street Bridge, New York City (foreground) and hinge
connection of that bridge (International Structural Slides, Berkeley, CA).
3.4.2

Example 3.6
40 kN
B

50 kN

1.00 m 1.00 m

3.00 m

40 kN

2.00 m

2.00 m

10 kNm
1.00 m 1.00 m

Fig. 3.23 Example 3.6.


Problem: Find the reaction forces at A,C and D of the beam for the given loading.
Solution: As discussed above, we cut the beam at the internal pin and apply the equations of equilibrium to the resulting two
free-body diagrams.

40 kN

FBD I
A

1.00 m 1.00 m

Bx

Bx

By

By

10 kNm

3.00 m

40 kN

50 kN

FBD II

2.00 m

2.00 m

Dy

Dx
1.00 m 1.00 m

Fig. 3.24 Free-body diagrams of two beam segments.


FBD I

F
M
F

=0

Bx = 0

= 0 = 40 1 + By 2

Bx = 20 kN

= 0 = 40 + By + A = 40 + 20 + A

A = 20 kN (ans )

=0

Dx = 0(ans )

= 0 = 20 3 50 2 40 5 10 + Dy 4

Dy = 62.5 kN (ans )

= 0 = 20 50 40 + 62.5 + C =

C = 47.5 kN (ans )

x ,I
A,I

y ,I

(3.21)

FBD II

F
M
F

x ,II

C ,II

y ,II

40 kN

40 kN

50 kN

10 kNm
47.5 kN

20 kN
1.00 m 1.00 m

3.00 m

2.00 m

62.5 kN
2.00 m

1.00 m 1.00 m

Fig. 3.25 Free-body diagram of whole structure with known reaction forces.
39

Problems
3.1

1.5 in

3 in

1.5 in

70D 500 lb
5 in

Find the moment about A of the force shown, (a) by using the definition of a moment (force x perpendicular distance),
(b) resolving the force into horizontal and vertical components and applying Varignons theorem.
Solution: M A = 362.7 lb-in (cw)
3.2

50 kN

100 kNm

1.5 m

10 kN B

1.5 m

30D
80 kN
A
5m

Find the moment of the applied forces and moments (a) about point A , (b) about point B .
(a) M A = 683.9 kNm

Solution: sign convention is ccw+:


3.3

1 kN 2 kN

(b) M B = 446.1kNm

4 kN 5 kN 6 kN 7 kN

3 kN

8 kN
240 kNm

1.00 m 1.00 m 1.00 m 1.00 m 1.00 m 1.00 m 1.00 m 1.00 m

Determine the magnitude of the resultant of the above force-moment system and its point of application.
3.4

(a)

F
F
r

(b)

(c)

F
F

F
r

For each of the three cases, replace the given system of forces by an equivalent force-moment system at A . If possible,
also replace the force-moment system by a single force. All applied forces have magnitude F .
40

70D

2 in

100 lb

2 in

3.5

2 in

(a) If the combined moment of the two forces about point A is zero, find the magnitude of the force F .
(b) What is the combined moment of the two forces about point O (use the value for F found in (a))
(c) Find the x - and y - coordinates of the two points, where the resultant of the two forces intersects the circle.
Solution: (a) F = 76.87 lb (b) MO = 222.1 lb-in (c) x 1 = 3.81 in y1 = 1.22 in, x 2 = 1.84 in y2 = 3.55 in
3.6

5k

6 ft

1k

5 k-ft
a/2

a/2

Find the required width a of the footing such that the line of action of the resultant of the given force-moment system
passes through the footing, i.e. passes through A .
Solution: a = 4.4 ft

5 ft

20 k

5 ft

10 k
5k

5 ft

30 k

5 ft

3.7

Find the magnitude of the moment M such that the resultant of the four forces and the moment applied to column
AB passes through the mid-height of the column.
Solution: M = 375 k-ft
41

1975 k-ft

5 ft

F2

100 k

5k
A

5 ft

5 ft

10 k

5 ft

5 ft

20 k

F1

5 ft

5 ft

30 k

5 ft

3.8

100 k
A

(a)
(b)
Find the magnitude of forces F1 and F2 such that the loading conditions in (a) and b) are equivalent.
Solution: F1 = 85 k, F2 = 50 k

3.9
(a)

(b)

(c)

(d)

(e)

(f)

For each of the six support conditions above, determine whether the body is statically determinate, unstable (improperly constrained), or statically indeterminate.
(a)

(b)

3.10
a

(c)

(d)

(e)

L/2

(f)

L/2

L/2

L/2

(g)

(h)

L/2

Determine the support reactions for given F , M , a, b, L .

42

F
L/2

F
2

MA

Solution:

a, b, c, e, g

Ax

Ay

Ax

Ay

(a)

(b)

b
a +b
M
L

a
a +b
M

(c)

F cos

F sin

(d)

(e)

(f)

(g)

(h)

F
2

3.11

d, f, h

Ax
Ay

MA

b
a +b

F sin

a
a +b
F

b
a

b
F (1 + )
a

b
a +b

L
2

b
a +b
0

0.40

[m]
B

60D
100 kN

0.30

Determine the horizontal and vertical support reactions under the given loading.
Solution: A = 37.5 kN , Bx = 124.1 kN , By = 50.0 kN
10 kNm

3.12

30D

6.00 m

30D

Determine the support reactions under the given loading.


Solution: B = 1.925 kN 3 , Ax = 0.963 kN , Ay = 1.667 kN

43

3.13

5k

24 ft

10 ft

10 ft

Assume that the support at A is unable to resist uplift, i.e. it works only in compression, such that uplift of the structure
from support A constitutes failure. Calculate the maximum allowable force F .
Solution: F = 2.08 k
M0

M0

3.14

(a) Calculate the support reactions for the given loading.


(b) Calculate the support reactions in absence of the right-hand moment M 0 .
Solution: (a) By = 0, Ay = 0, Ax = 0

(b) By =

M0
M
, Ay = 0 , Ax = 0
L
L

3.15
80 kN

100 kN

200 kNm

80 kN

C
3.0 m

B 100 kN

4.0 m

4.0 m

200 kNm

C
3.0 m

3.0 m

Find the support reactions for the two structures shown.


44

3.0 m

3.16

5 kN

0.50 m

2.25 m

0.75 m

Two loads act on a beam, which is supported by two cables attached at A and B . Neglecting the weight of the beam,
find the range of values of F for which neither cable becomes slack.
Solution: 0.833 kN < F < 18.33 kN
3.17

B
W

5m

A person of weightW climbs up a ladder placed between points A and B . The maximum friction force (the horizontal
force) developed at point A is 20% of the vertical force at A . Ignore friction at point B .
(a) For = 45D find the distance x at which the ladder starts to slide,
(b) Find the required angle such that the ladder is safe regardless of the location x of the weight.
Solution: (a) x = 1m, (b) = 78.7 D
3.18
(a)

20 kN
A

1.00 m

40 kN

3.00 m

1.00 m

20 kN

1.00 m

2.00 m

20 kN

(b) A

2.00 m

2.00 m

2.00 m

2.00 m

1.50 m

Find the reaction forces of the beams for the given loading.

45

20 kN
D

1.00 m

3.00 m

15 kNm

1.50 m

1.00 m

Solution:
(a) Ay = By = C y = Dy = 20 kN , Ax = 0
(b) Ay = 5 kN , B = 30 kN , C = 5 kN , Ax = 0, M A = 5 kNm (cw)

3.19

120 kN
F

4.00 m

2.00 m

50 kN

C
1.00 m

3.00 m

3.00 m

1.00 m

(a) For F =200 kN calculate the reactions at supports A, B, and C .


(b) Find the range of values of F such that both the supports at A and B react in compression.

100 kN

3.20

60D

B
4m

Find the reaction forces at supports A and B of the three-pinned arch for the given loading.

46

4. Statics of objects in three dimensions


4.1

Moments of a 3-D force

4.1.1

Discussion

In Chapter 3, we introduced the moment of a force about a point as the product of force and perpendicular distance. For
two-dimensional structures lying in the x -y plane the moment is about the z - axis, i.e. an axis perpendicular to the drawing
plane. For three-dimensional structures, moments exist about all three axes in space, i.e. M x , M y and M z .
F

z
y

x
Fz
Fy

Fx

dz
dx

dy

Fig. 4.1 Moment of force F acting at B about point A .


When calculating moments in three dimensions, we use Varignons principle and proceed as follows (see Fig. 4.1):
(1) resolve the vector AB from the moment reference point to the point of application of the force into its three rectangular
components dx , dy and dz .
(2) resolve the force vector into its three rectangular components Fx , Fy and Fz
M y = dz Fx

M z = dy Fx

M x = dz Fy

M x = dy Fz

M z = dx Fy

M y = dx Fz

Fz

Fx
Fy

dz
dx

dy
A

B
dz

dx

dy

dz
dx

dy

Fig. 4.2 Calculating a moment of a force in three dimensions using rectangular components.

47

Moment due to Fx
The perpendicular distances from the reference point A to the line of action of the force Fx are dy and dz . The tendency of the
force component Fx to rotate about A is (1) about the y - axis through moment arm dz and (2) about the z - axis through
moment arm dy . We note that the rotation about the y - axis is positive, the rotation about the z - axis, however, is negative.
Hence the moment caused by Fx is
M y (due to Fx ) = dz Fx

M z (due to Fx ) = dy Fx

(4.1)

Moment due to Fy
The perpendicular distances from the reference point A to the line of action of the force Fy are dx and dz . The tendency of the
force component Fy to rotate about A is (1) about the x - axis through moment arm dz and (2) about the z - axis through
moment arm dx . We note that the rotation about the z - axis is positive, the rotation about the x - axis, however, is negative.
Hence the moment caused by Fy is
M x (due to Fy ) = dz Fy

M z (due to Fy ) = dx Fy

(4.2)

Moment due to Fz
The perpendicular distances from the reference point A to the line of action of the force Fz are dx and dy . The tendency of the
force component Fz to rotate about A is (1) about the x - axis through moment arm dy and (2) about the y - axis through
moment arm dx . We note that the rotation about the x - axis is positive, the rotation about the y - axis, however, is negative.
Hence the moment caused by Fy is
M x (due to Fz ) = dy Fz

M y (due to Fz ) = dx Fz

(4.3)

We summarize:
A moment about the x - axis is caused by a force along the y - axis and a moment arm in the z - direction
or a force along the z - axis and a moment arm in the y - direction
M x = dy Fz dz Fy

(4.4)

A moment about the y - axis is caused by a force along the x - axis and a moment arm in the z - direction
or a force along the z - axis and a moment arm in the x - direction
M y = dz Fx dx Fz

(4.5)

A moment about the z - axis is caused by a force along the x - axis and a moment arm in the y - direction
or a force along the y - axis and a moment arm in the x - direction
M z = dx Fy dy Fx

(4.6)

z, M z
y, M y
x, M x

Fig. 4.3 Sign convention for moment.


Important: In order to calculate the correct sign of the moment, we need to calculate the distances dx , dy , dz as components
of the vector from the reference point A about which we take the moment to the point of application B of the forces (not
from the point of application to the reference point), hence (see Fig. 4.1)
dx = x B x A

dy = y B y A

dz = z B z A

(4.7)

48

M x = dy Fz dz Fy

M y = dz Fx dx Fz

Fz

Fz

Fy

Fx
dz

dy

M z = dx Fy dy Fx

dx

dz
dy

Fx

Fy

dx

dz
dy

dx
A

Fig. 4.4 Moment about three axes passing through a point A .


4.1.2
Example 4.1
Problem: A force F with a magnitude of 10kN acts at origin O . The line of action of F passes through point A , whose coordinates are x =5 m, y =4 m, z =3 m . Find the moment of the force F about point B whose coordinates are x =8 m ,
y=7 m and z=6 m
Solution:
We first resolve the force into its components along the three coordinate axes. The distances between points O and A along
the three coordinate axes are
x = x A x o = 5 0 = 5 m

y = yA yo = 4 0 = 4 m

z = z A zo = 3 0 = 3 m

(4.8)

such that
L = x 2 + y 2 + z 2 = 52 + 42 + 32 = 7.071 m

(4.9)

The three rectangular components are thus


Fx = 10

5
4
= 7.071 kN Fy = 10
= 5.657 kN
7.071
7.071

Fz = 10

3
= 4.243 kN
7.071

(4.10)

Next, we find the distances between the moment reference point B and the point of application O of the force. Using Eq.
(4.7) gives
d x = xO x B = 0 8 = 8 m

dy = yO yB = 0 7 = 7 m

dz = zO z B = 0 6 = 6 m

(4.11)

Using Eqs. (4.4-4.6), we calculate the three components of the moment


M x = dy Fz dz Fy = 7 4.243 + 6 5.657 = 4.24 kNm (ans )
M y = dz Fx dx Fz = 6 7.071 + 8 4.243 = 8.48 kNm (ans )

(4.12)

M z = dx Fy dy Fx = 8 5.657 + 7 7.071 = 4.24 kNm (ans )

Note:
If we move the force along its line of action, i.e. consider it applied at a different point, the moment of that force about
point B does not change. For example if we consider the force applied at A instead of O , the components of the vector
from B to A become
d x = x A x B = 5 8 = 3 m

dy = yA yB = 4 7 = 3 m

49

d z = z A z B = 3 6 = 3 m

(4.13)

and the moment is equal to


M x = dy Fz dz Fy = 3 4.243 + 3 5.657 = 4.24 kNm (ans )
M y = dz Fx dx Fz = 3 7.071 + 3 4.243 = 8.48 kNm (ans )

(4.14)

M z = dx Fy dy Fx = 3 5.657 + 3 7.071 = 4.24 kNm (ans )

a result identical to that above.

4.2

Resultant of parallel forces in 3-D

4.2.1

Discussion

In Chapter 3, we have learned how to replace a given set of forces applied to a two-dimensional body by (1) an equivalent
force at an arbitrary location and a moment or (2) by an equivalent force (and no moment) at a specific location. We can use
an analogous procedure in three dimensions. A special case exists when all forces applied to a body have the same direction,
say along the z -axis. We can then replace at an arbitrary location the applied forces by an equivalent resultant force, a
moment about the x -axis and a moment about the y -axis. Note that since all forces are applied in the z -direction, the
resultant forces in the x -and y -directions and the moment about the z -direction are zero. This procedure is analogous to
method (1) in 2-D. We can also calculate a specific location where we have to apply the resultant force such that there are
no resultant moments. This procedure is analogous to method (2) in 2-D.
4.2.2

4k

Example 4.2
2k

8k
6k

2 ft

y
O

3 ft

Fig. 4.5 Parallel forces in 3-D (Example 4.2).


Problem: (a) Replace the above four forces by their resultant force at O and their resultant moment, (b) Find the location
where we can replace the above four forces by their resultant force (zero moment).
Solution: (a) The resultant force of the four applied force above is
Rz = + Fz = 2 4 6 8 = 20 k (ans )

(4.15)

Taking moments about O gives

O ,x

= 2 ft 2 k 2 ft 4 k = 12 k-ft(ans )

O ,y

= 3 ft 4 k + 3 ft 8 k = 36 k-ft(ans )

(4.16)

(b) As in 2-D, we move the resultant force away from point O until the moments become zero. Since the moments about
both the x -and y -axes need to vanish, we have to move the force in both the x -and y -directions. Hence
xc = 36 k-ft/20 k = 1.8 ft

yc = 12 k-ft/20 k = 0.6 ft

(4.17)

We will return to the concept of finding the location of the resultant of parallel forces in 3-D in Chapter 5 when we discuss
the topic of center of gravity. Fig. 4.6 summarizes the results of this example.

50

4k

20 k
8k

2k
6k

2 ft

12 k-ft

20 k

3 ft

z 36 k-ft

0.6 ft

z
x

1.8 ft

Fig. 4.6 Resultant of parallel forces in 3-D.

4.3

Equilibrium of parallel forces in 3-D

4.3.1

Discussion

For a structure subject to a set of parallel forces, say in the z -direction, we can write three equations of equilibrium

=0

=0

=0

(4.18)

Since no forces act in the x -and y -directions, the remaining three equations of equilibrium (2 force equations and one moment equation) are automatically satisfied. Note that the most common source of parallel forces is gravity.
4.3.2

Example 4.3
F

z
y

1
2

2
10 ft

Fig. 4.7 Example 4.3.


Problem: Three vertical columns spaced at 120 degrees support a circular foundation mat of 10 ft radius against a vertical
force of magnitude F = 10 k . Diagonal members stabilize the structure for potential lateral loading which is absent in this
problem (diagonal forces are zero). Find the forces in the three columns.
Solution: We calculate the three unknown forces using the equations of equilibrium in Eq. (4.18). As always, the first step is
to draw a free-body diagram. As already mentioned in the previous 3-D example, we can draw a three-dimensional freebody diagram and/or consider projections onto several planes (side and front views, i.e. x -z and y -z projections) shown Fig.
4.8. The sense of the three column forces assumed on the free-body diagram is compression. For the equation that sums
moments about the x -axis, we consider the projection onto the y -z plane, for the equation that sums moments about the y axis, we consider a projection onto the x -z plane. We can use either free-body diagram for the equation that sums forces in
the z -direction.

51

10 k
z
y

F1
1

F2

O
F3

5 ft

F2

y
F3

8.660 ft

10 ft

10 k

F1
8.660 ft
10 ft

10 k
z

F1

F3

Fig. 4.8 Free-body diagrams (projections and 3-D).


The three equations of equilibrium are

F = 0
M = 0
M = 0

F2

= F1 + F2 + F3 10

O ,y

= F1 5 + F2 5 F3 10

O ,x

= F1 8.660 F2 8.660 10 10

(4.19)

Solving the set of three simultaneous equations, gives


F1 = 9.11 k

F2 = 2.44 k

F3 = 3.33 k

(ans )

(4.20)

Note that we can use any reference point for the moment equations. In fact when solving the equations in Eq. (4.19)
manually without pocket calculator, it is advantageous to pick the point of application of forces F1 and F2 as the reference
point for the first moment equation, which allows us to solve for F3 directly.

=0

= 10 5 F3 15 F3 = 3.33 k

(4.21)

Knowing F3 we can then solve sequentially for F1 and F2 .


The negative numerical value obtained for F2 indicates that the force in column 2 is a tension force. Columns 1 and 3 react
in compression as assumed on the free-body diagrams in Fig. 4.8. As a summary of the results, we draw a free-body
diagram with all forces known. Note that we have reversed the directions of the arrow representing the force in column 2
reflecting the negative value obtained for F2 in Eq. (4.20).

10 k
z

y
9.11 k

x
3.33 k

2.44 k

Fig. 4.9 Summary of results (columns 1 and 3 in compression, column 2 in tension).

52

4.4

Equilibrium of general forces in 3-D

4.4.1

Discussion

In order to solve a 3-D equilibrium problem involving forces in arbitrary direction, we need to use all six equations of
equilibrium

=0

4.4.2

=0

=0

=0

=0

=0

(4.22)

Example 4.4

4 kN
1m

C
6
1

1m

2
3

1m

3m

Fig. 4.10 Example 4.4.


Problem: The box above is stabilized by six links and loaded as shown. Find the forces in the links. Clearly indicate
whether the links are in tension or compression.
Solution:
We calculate the six unknown forces using the equations of equilibrium in Eq. (4.22). Like in the previous example, the first
step is to identify correctly all forces acting on the body by cutting through the links and drawing the free-body diagram.
Fig. 4.11 shows a free-body diagram as a 3-D view. As always, the direction we select for the unknown forces in the links is
arbitrary. Here, we assume a tension force in all links (force is directed away from the body).
z
4 kN
4 kN
y
x
1m
1m
B

C
6
1
1m

1m

4
1m

1m

2
3

3m

C
L6

L5

A
L2

L1

L4

3m

L3

Fig. 4.11 Free-body diagram (3-D view).


As in the previous examples, we can write the equations of equilibrium more easily if we draw different projections (views)
of the structure. Since we now use all three moment equations, we need three different projections.
(1) When taking moments about the x -axis, we consider a projection onto the y -z plane.
(2) When taking moments about the y -axis, we consider a projection onto the x -z plane.
(3) When taking moments about the z -axis, we consider a projection onto the x -y plane.

53

3m

3m

L2

L3

L4

L6

L5

L2

L4

A, B

L3

L5

1m

L1 A, C

1m

L1 A

2m

2m

L6
y

Fig. 4.12 Free-body diagrams (projections).


From the x -y projection we can write the following three equations of equilibrium.

F = 0
M = 0
F = 0
x

A,z

= 4 L1

L1 = 4 kN

(1)

= 4 1 L4 3

L4 = 1.333 kN

(2)

= L2 L4

L2 = L4 = 1.333 kN

(3)

(4.23)

From the x -z projection we can write the following two equations of equilibrium (we have used the third equation already)

M = 0
F = 0
A,y

= 4 1 + L5 3

L5 = 1.333 kN

= L3 L5 L6

(4)
(5)

(4.24)

From the y -z projection we can write the final equation of equilibrium.

B ,x

= 0 = L6 2

L6 = 0

(6)

(4.25)

Substituting the results of equations (4) and (6) into (5), gives
L3 = 1.333 kN

(4.26)

The negative numerical values for L1, L2 , L3 indicate that the corresponding forces act in the direction opposite to our
assumption, i.e. in compression. As a summary of the results, we draw a free-body diagram with all forces known.
4 kN

1m
1.333 kN
1m

0
1 m 4 kN

1.333 kN
1.333 kN
3m

1.333 kN

Fig. 4.13 Summary of results (Links 1,2,3 in compression, links 4,5 in tension, link 6 is zero-force link).

54

4.4.3

Example 4.5

5m
2

z
1

10 kN

20 kN
5m
5m

3m

4m

Fig. 4.14 Example 4.5.


Problem: Line AB is pinned to the ground at A , supported by two cables and loaded as shown. Find the tension forces in
the two cables.
Solution: We first express the three components of the cable forces in terms of their unknown magnitudes F1 and F2
L1 = 32 + 42 + 102 = 125

L2 = 52 + 102 = 125

4
F1 = 0.3578 F1
125

F1x =
F2x = 0

F1y =

3
F1 = 0.2683 F1
125

F1z =

10
F2 = 0.8944 F2
125

F2y =

5
F2 = 0.4472 F2
125

F2z =

10
F2 = 0.8944 F2
125

(4.27)

Recognizing that the two cables prevent the line from rotating about A , the best solution strategy is to write two moment
equations with point A as the reference point. Furthermore, we observe that cable 1 is the only cable resisting a rotation
about the y -axis. Thus it is a good idea to start with summing moments about the y -axis, which gives us an equation that
only involves one unknown.

Ay

= 0 = 10 5 F1x 10 = 50 0.3578 F1 10

F1 = 13.97 kN(ans )

(4.28)

Summing moments about the x -axis then gives

Ax

= 0 = F1y 10 F2y 10 + 20 5 = 13.97 0.2638 10 0.4472 F2 10 + 100 F2 = 30.60 kN(ans )


B
F2

5m
2

z
1

10 kN

F1

10 kN

20 kN

20 kN

5m
Ay

5m

Ax

3m

Az

4m

Fig. 4.15 Free-body diagram.

55

(4.29)

Problems
z

4.1
y
B

F = 100 lb

10 in

30 in

20 in
O

Calculate the three rectangular components M x , M y , M z of the moment of the force F about point O . (a) Consider the
force acting at A . (b) Consider the force acting at B . Show that the results are identical.
4.2

(d)
(c)
(b)
(a)
(h)

(g)

x
(f)

(e)

(l)
(k)
(j)
(i)
(p)
(o)
(n)
(m)

Each of the prisms above has dimensions of 1, 2, 3 along the x - , y - and z - direction respectively. All applied forces
have unit magnitude. Find the moments of the force applied to each prism about the x - , y - and z - axes passing through
point A , the bottom-left-front corner of the prisms below.
56

4.3
2k

2k

2k

3 ft

3 ft

3 ft

(c)

(b)

(a)
5 ft

5 ft

4 ft

5 ft

4 ft

For the three loads cases shown, find the moments about the x - , y - and z - axes passing through point A .
4.4
50 kN
75 kN
20 kN

50 kN
75 kN

3m

z
y

30 kN

2.50 m

3m
2.50 m

3m

Find the moments about the x - , y - and z - axes passing through point O .
4.5
30 kN
50 kN

75 kN
25 kN

2.50 m

3m
2.50 m

3m

Determine the magnitude of the resultant of the above forces and its point of application.
Solution: R = 180 kN (downward), x R = 0.333 m, yR = 0.417 m

57

4 ft

4.6

F2

F7

F1

F5

F4

F3

F6

12 ft

F
2

16 ft
D

G
9 ft

9 ft

16 ft

9 ft
Find the forces in the five links for the given loading (consider the seven forces separately). All applied forces have
10k magnitude. Clearly indicate whether the links are in tension or compression.
Solution:
Link
F1
F2
F3
F4
F5
1
0
0 13.33
3.75
5
2
0
0
16.67
0
0
3
0
0
0
6.25
0
4
7.5
10
0
3.75
5
5
12.5
0
0
6.25
0
Note: positive = tension, negative = compression, forces in kips
4.7

F6

F6
7.5
0
12.5
0
0

F7
F5
F3

F8
F9

F4

12 ft

F7
10
0
0
0
0

F1

F2

z
y

4
5
12 ft
16 ft
9 ft
7 ft

12 ft

Find the forces in the six links for the given loading (consider the nine forces separately). All applied forces have 10k
magnitude. Clearly indicate whether the links are in tension or compression.
Solution:
Link
F1
F2
F3
F4
F5
F6
F7
F8
F9
1
10.00
7.50
10.00
0
15.63
0
10.00
7.50 15.63
2
0
0
0 13.33 10.00
13.33 10.00
0
10.00
58

3
0
0
0
0
0
0
4
14.14
0 14.14
0 14.14
0
5
0
12.50
0
0
9.38
0
6
0
0
0
16.67
12.50 16.67
Note: positive = tension, negative = compression, forces in kips.

10.00
0
0
0

0
0
12.50
0

0
14.14
9.38
12.50

4.8

F4

F5

F3
F1

F7

1m

F6

C
6
1

2m

F2
4

A
2

3m

Find the forces in the six links (consider the seven forces separately). All applied forces have 10 kN magnitude.
Solution:
Link
F1
F2
F3
F4
F5
F6
F7
1
10.00
0
10.00
0
0
10.0
0
2
0
0
6.67
0
0
6.67
10.0
3
3.33
5.00
3.33
5.00
10.00
3.33
5.00
4
0
10.00
6.67
10.00
0
6.67
0
5
3.33
0
3.33
0
10.00
3.33
0
6
0
5.00
0
5.00
10.00
0
5.00
Note: positive = tension, negative = compression, forces in kN.

59

5. Distributed loads, center of gravity, centroid


5.1

Introduction

A
L

So far, we have considered all forces as concentrated at their points of application. Strictly speaking, concentrated forces do
not exist since every applied force is distributed over a finite area even though the area can often be small. One of the most
obvious phenomena in the context of distributed load is self-weight of a structural component. Shown in Fig. 5.1 is a beam
with linearly varying depth d(x ) , constant thickness b (dimension out of plane) and density .
density
variable depth
constant thicknessb

Fig. 5.1 Beam of variable depth.


Lets assume for a moment that the self-weight of the beam is the only force applied to it and that we are interested in the
support reactions at A . We observe that the gravity forces, the weight of the beam, act as a distributed load. Since the depth
of the beam varies along its length, the intensity of the distributed load under gravity g , expressed in force per length is
w(x ) = g b d (x )

[force/length]
w(x )

Fig. 5.2 Distributed load due to self-weight.


Since we have already learned how to calculate the support reactions due to concentrated forces, we wish to replace the distributed load w(x ) by its resultant concentrated forceW . We observe that the resultant forceW is equal to the total weight
of the beam. The total weight in turn is equal to the area formed by the intensity w(x ) and the length L over which the load is
distributed, hence
L

W =

w(x )dx

(5.1)

The preceding equation gives us the magnitude of the resultant. The more difficult question that arises in dealing with distributed loads is the location of the resultant. Recall from our discussion in Chapter 3 that we calculate the resultant based on
the condition that it must have the same effect (regarding translation and rotation) on the structure as the original forces
acting on it.
To shed some more light on the phenomenon of distributed loads and where to replace the distributed load by a concentrated force, we continue this introductory section with a small experiment. We take a triangle, put it on a desk and slowly
slide it toward the edge of the desk. We notice that although part of the triangle is overhanging, the triangle still remains in
its stable horizontal position on the desk. When the overhanging portion of the triangle reaches a certain amount, the triangle overturns and falls off the table. Obviously, this critical distance has a special significance. We call the position on the
triangle defined by the critical distance the center of gravity of the triangle.

60

xc

Fig. 5.3 Triangle falls when resultant leaves the table.

5.2

Center of gravity, center of mass, centroid

5.2.1
Theory-principle of moments
We now turn to the question of how we can determine the center of gravity mathematically. We revisit the example of a
triangular plate of constant thickness b .
xc

d
x
L

dx

Fig. 5.4 Centroid of triangle.


Our goal is to replace the distributed triangular load by a single concentrated force at the appropriate location which is the
center of gravity. The magnitude of the concentrated force is the volume of the body multiplied by its density and
gravitational constant g
L

W =

w(x )dx = g b d (x ) dx = g b
0

d
1
x dx = g b d L
L
2

(5.2)

Note that for the simple triangle above formal integration is unnecessary, since we know the area of a triangle and the total
weight is simply area x thickness x density x gravitational constant. The question of where the concentrated force W acts is
more complicated. Since we want the concentrated forceW to have the same tendency to rotate about any point as the
distributed load, we require that the moment of the total weightW about any axis equals the sum of the moments of all
small elements dW about the same axis. We select the origin as the reference point. The moment of the total weightW is
M = xc W

(5.3)

where xc is the unknown coordinate of the center of gravity. The sum of the moments of all small elements dW is
L

M =

x w(x ) dx = g b x d (x ) dx = g b
0

d 2
1
x dx = g b d L2
L
3

(5.4)

Setting the two equal gives us an equation for the location of the center of gravity
M
xc =
=
W

x dW
dW

1
g b dL2
2
3
=
= L
1
3
g b dL
2

(5.5)

61

Note that in our experiment we are interested in finding the x - coordinate of the center of gravity. With analogous expressions for the coordinate yc , we have
yc =

y dW

(5.6)

By inspection, we obtain
=

yc

1
h
3

(5.7)

When the density and the gravitational constant g as well as the thickness are constant throughout the body, they will be a
constant factor in both the numerators and denominators of Eqs. (5.5) and (5.6) and cancel. The remaining equations define
a purely geometrical property of the body, because both mass and gravity have disappeared. When our calculations concern
the geometry of the body only, we use the term centroid. Equations 5.5 and 5.6 then become
xc =

x dA = x dA
A
dA

yc =

y dA = y dA
A
dA

(5.8)

We call the numerator in the preceding equation the first moment of the area A , since the respective moment arms x and y
appear with exponent of one. Higher moments of areas, e.g. x 2dA , also play an important role in mechanis but are beyond the treatment in this class.
We summarize: The center of gravity is the location at which we can replace the distributed weight of a body by its total
weight. Since the gravity field is constant for all practical purposes, the center of gravity coincides with the center of mass.
Moreover, if the density of the material is uniform throughout the body, the center of mass coincides with the centroid of the
volume. If, in addition the thickness of the body is uniform the centroid of the body coincides with the centroid of the area
that forms the body, in our case the triangle. It is important to recognize that the centroid is a purely geometrical property of
a body since it does not involve the mass or gravity. In the remainder of this chapter, we are concerned with calculating
centroids of areas.
5.2.2

Example 5.1

w2
w1
x
L

Fig. 5.5 Centroid of trapezoid.


Problem: Find the x - coordinate of the centroid of the above trapezoid.
Solution: According to Eq. (5.8), we need to calculate the area A= dA and the first moment x dA of the trapezoid (if
we asked for the y - coordinate of the centroid, we would also have to determine the first moment y dA ). Finding the integrals dA and x dA is different from a conventional math integration problem, because the function that we need to integrate and the integration limits are not readily given to us. The challenge is thus to find the integrand along with the limits
of integration.
62

In this example, we need to express the height of the trapezoid as a function of x . Recognizing that the top side of the trapezoid is a straight line with slope (w2 w1 )/ L and intercept w1 , we write (see Fig. 5.6)
y

w2
w1

dA = y(x ) dx

y(x )

dx
L

Fig. 5.6 Differential element.


L

A=

dA =
0

w w1

y(x ) dx = w1 + 2
x

L
0

w 2 w1 x 2

dx = w1x +
2
L

= w1L + (w2 w1 )

L w1 + w 2
=
L
2
2

(5.9)

Note that we could have also arrived at the above result without formal integration by calculating the area of the trapezoid
as the average height (w1 +w2 )/2 times the length L .
The second integral we have to calculate is the first moment of the area A (sum of the moments of the differential
areas dA ). Since we already have the expression for dA , all we need to do is multiply by the moment arm x and obtain
L

x dA

x y(x ) dx =
0

w w1

x w1 + 2
x

x 2 w 2 w1 x 3

+
dx = w1
2
3
L

(5.10)

L2
L2
L2
= w1
+ (w 2 w1 )
=
(w1 + 2w2 )
2
3
6

The x - coordinate of the centroid is thus


xc =

x dA
A

L2
(w1 + 2w2 )
L w1 + 2w2
= 6 w +w
=
3 w1 + w 2
2
L 1
2

(5.11)

Note the two special cases


L w + 2w
L
=
3 w +w
2

w1 = w 2 = w

xc =

w1 = 0; w2 = w

L 2w
2
xc =
= L
3 w
3

rectangle

(5.12)
triangle

63

5.2.3

Example 5.2

x = y 2/2

y = x 3/4

2
Fig. 5.7 Area defined by two curves.
Problem: Find the coordinates of the centroid of the shaded area between the two curves in Fig. 5.7.
Solution I: We select x as the variable for integration. As in the previous example, we need to express the area of differential
element in terms of x . We notice that the top curve is defined as x = f (y ) . To find y = f (x ) for the top curve, we first need to
find the inverse
x =

y2
y = 2x
2

(5.13)

Hence the area of the differential element dA is (see Fig. 5.8)


x3

dA = 2x dx
4

(5.14)

We first find the area by evaluating the integral


2

A=

dA =

x3
2 3
1 4
2x dx = 2 x 2 x
4
3
16

= 1.667

(5.15)

Like in Example 5.1, we simply multiply by x the function we integrated in the preceding equation to find the first moment
of the area about the y - axis and obtain
2

x dA =

x3
2 5
1
x 2x dx = 2 x 2 x 5 = 1.600

4
5
20 0

(5.16)

The x - coordinate of the centroid of the shaded area is thus


xc =

1.600
= 0.960
1.667

(5.17)

64

y = 2x

y = x 3/4

dx
2

Fig. 5.8 Differential element for integration along x .


Now we have to find yc , the y - coordinate of the centroid. Caution: Since the differential element has finite length in
the y - direction, the procedure becomes somewhat more complicated since we first have to locate the centroid of the differential element. Since the differential element is a rectangle, the y - coordinate of the centroid is
y =

1
x3
2x +
2
4

(5.18)

such that
2

y dA =

2
1
x 3
x 3
1

x
x
x
2
2
d
+


2
4

4
2 0


y

2
6

2x x dx = 1 x 2 1 x 7 = 1.429
16
2
112 0

(5.19)

dA

The y - coordinate of the centroid of the shaded area is thus


yc =

1.429
= 0.857
1.667

(5.20)

Solution II:

x = y 2/2

dy

x =(4y )3
x

2
Fig. 5.9 Differential element for integration along y .
We select y as the variable for integration and need to express the area of the horizontal differential element in terms of y .
Since the bottom curve is defined as y = f (x ) , we find the inverse
65

y=

1
x3
x = ( 4y ) 3
4

(5.21)

We first find the area by evaluating the integral


2

A=

dA =

2
1
4
1

(4y )3 y dy = 4 3 3 y 3 1 y 3

2
4
6

= 1.667

(5.22)

We next find the first moment of the area about the y - axis by calculating
2

x dA =
0

2
1
1
1
y2
y2
1
(4y )3 + (4y )3 dy =
2
2

2
2 0



x

2 2 y 4
1 2 3 5 y5
4 3 y 3 dy = 4 3 y 3 = 1.600
4
2 5
20 0

(5.23)

dA

In manipulating the integrand, we have used the identity

(a + b ) (a b ) = a 2 b 2

(5.24)

Noting that y =y for the horizontal element, we find the first moment of the area about the x - axis by calculating
2

y dA =

1
1

y2
3 7 1
y (4y )3 dy = 4 3 y 3 y 4

2
7
8

= 1.429

(5.25)

The results are the same as those of Solution I.

5.3

Calculating centroids of composite areas

5.3.1
Discussion
The majority of engineers and architects are probably somewhat allergic to the formal integration process discussed in the
previous section. Fortunately, there is often an easier procedure to calculate the centroid. In many applications, we can conveniently subdivide an area A of interest into rectangles, triangles, circles or other common shapes Ai whose individual centroids we know. We can determine the centroid of the composite area A from the centroids of the various parts Ai by expressing that the moment of the composite area is equal to the sum of the moments of the individual areas. We assume that
the area under consideration is the sum of individual areas Ai
n

A = Ai

(5.26)

The moment of the whole area is then


M y = xc A

(5.27)

and

M x = yc A

(5.28)

where xc and yc are the unknown coordinates of the centroid of the composite area. The sums of the moments of the individual parts are
M y = x c1A1 + xc 2A2 + ... + x cn An

(5.29)

and
M x = yc1A1 + yc 2A2 + ... + ycn An

(5.30)

where xci and yci are coordinates of the centroid of the individual parts of the area. Hence
n

xc =

xc1A1 + x c 2A2 + ... + xcn An


A

x
i =1

ci

Ai
(5.31)

A
66

and
n

yc =

yc1A1 + yc 2A2 + ... + ycn An


A

i =1

ci

Ai
(5.32)

A3

0.50

Example 5.3

A2

1.00

5.3.2

A1

x
0.50

2.00

Fig. 5.10 Composite area.


Problem: Calculate the location of the centroid for the area in Fig. 5.10.
Solution: The area A is composed of a rectangle A1 and two triangles A2 and A3 . Since we know the location of the centroids
of A1, A2 and A3 , we can use Eqs. (5.31) and (5.32) to calculate the centroid of A . First, we find the area
A = 1.0 2.0 +

1
1
0.5 2.0 + 0.5 1.5 = 2.0 + 0.5 + 0.375 = 2.875
2
2

(5.33)

Next, we calculate the sum of the products of the individual areas and the corresponding x - coordinate of their centroids

ci

Ai

1
1

= 0.5 + 2.0 2.0


+ 0.5 + 2.0 0.5
N + 0.5 0.375
N = 3.0000 + 0.5833 + 0.1250 = 3.708

N
2
3
3
A
A

1

2

A3
xc 1

xc 2

(5.34)

xc 3

The x - coordinate of the centroid of A is thus

xc =

3.708
= 1.29
2.875

Performing the same calculations for the y -direction, gives

ci

yc

1
Ai = 0.5 2.0 + 1.0 + 0.5 0.5 + 1.5 0.375 = 1.0000 + 0.5833 + 0.3750 = 1.9583

(5.35)

1.958
=
= 0.681
2.875

We have calculated the values for xc and yc with respect to the coordinate system in Fig. 5.10. For any other selection of the
origin we obtain different numerical results defining, of course, the same location.
When determining centroids of composite areas, it is advantageous to arrange our calculations in tabular form. The following table contains the calculations for the example above.

67

Table 5.1 Centroid of composite area calculations


i
Ai
xci
1
2.0000
1.5000
2
0.5000
1.1667
3
0.3750
0.3333
2.8750

xci Ai
3.0000
0.5833
0.1250
3.7080

yci
0.5000
1.1667
1.0000

yci Ai
1.0000
0.5833
0.3750
1.9583

If the composite area contains holes of any shape that we know the centroid of, we simply treat the holes as negative areas.
We illustrate the procedure by the example below.
Example 5.4

50

10

5.3.3

[in]

10

yc

50

Fig. 5.11 Centroid of T-section with holes.


Problem: Calculate the location of the centroid for the area in Fig. 5.11.
Solution: We use the tabular arrangement introduced above and account for the negative contribution of the two circular
holes. We consider only one half of the symmetrical area. Areas A1, A2, and A3 are the 10x50-rectangle (web), the 60x20-rectangle (flange) and the circular whole, respectively. Because of symmetry we can locate the horizontal position of the centroid by inspection (on the line of symmetry). Using the bottom of the section as our reference (origin of the y -axis) we
have
Table 5.2 Centroid of T-section
i
Ai [in2 ]
1
500
2
1200
3
78.54
1621.5

yi [in]
25
60
60

yi Ai [in 3 ]
12500
72000
4712
79788

The centroid is thus located at


yc =

79788[in 3]
1621.5[in 2 ]

= 49.2 in

(5.36)

68

5.4

Distributed load on beams

5.4.1
Discussion and introductory example (Example 5.5)
We can use the idea of the centroid of an area to calculate reactions for beams due to distributed loads by replacing the
distributed load on a beam by a concentrated force. The magnitude of the force is equal to the area under the load and the
location of its line of action coincides with the centroid of the area.
10 kN/m
8 kN/m
4 kN/m
2 kN/m
A

[m]

4.00

4.00

Fig. 5.12 Simply supported beam under trapezoidal loads.


Problem: A beam supports a distributed load as shown. Calculate the equivalent concentrated force and its location and determine the support reactions.
Solution: (1) Equivalent concentrated force and its location: The concentrated force is equal to the area under the load and
the line of action passes through the centroid of that area. We divide the area under the load into four triangles and construct
the table below to locate the centroid.
2 kN/m

8 kN/m
2

4
10 kN/m

3
1

4 kN/m

4.00 m

4.00 m

Fig. 5.13 Subdivision of load into triangles.


Table 5.3 Centroid of area under load
i
Ai

xci [m]

xci Ai

1
2
3
4

1.333
2.666
5.333
6.666

10.667
42.667
106.67
26.667
186.67

8
16
20
4
48

Thus the centroid is located at


xc =

186.67
= 3.889 m
48

(5.37)
69

3.88

48 kN
A

B
8.00

Fig. 5.14 Equivalent concentrated force.


(2) Support reactions: We have replaced the distributed load by a concentrated force and are now able to place the concentrated force on the free-body diagram of the beam along with the unknown support forces to apply the equilibrium equations. Note that after replacing the distributed load by an equivalent concentrated force, the problem has become a standard
equilibrium problem like those considered in Chapter 3.
3.89 m

48 kN
A

B
8.00 m

Fig. 5.15 Free-body diagram to calculate reactions.


Equilibrium yields

M A = 0 = B 8.00 48 3.889 B = 23.33


A = 24.67
Fy = 0 = 23.33 + A 48

(5.38)

Note that it is generally unnecessary to replace a distributed load by a single concentrated force. For example, we can use
four concentrated forces by replacing each of the four triangles by their own force. In this case, the principle of moments
that gave us the location of the centroid in Table 5.3 is part of the moment equation in Eq. (5.39).
6.66 m
5.33 m
2.66 m
1.33 m

8 kN

16 kN

20 kN

4 kN

8.00 m

Fig. 5.16 Alternate approach.


Equilibrium yields

MA = 0

= B 8.00 8 1.333 16 2.666 20 5.333 4 6.666


.67

= B 8.00 186


B = 23.33

see Table 5.3

Fy = 0 = 23.33 + A 48

A = 24.67

70

(5.39)

Example 5.6

10 k

10 k

2.00 ft

10 k/ft

1.50 ft

30D

Ay

1.50 ft

10 k/ft

1.50 ft

Ax

1.50 ft

5.4.2

20 k/ft

30D

2.00 ft

2.00 ft

20 k/ft

2.00 ft

Fig. 5.17 Example 5.6.


Problem: Determine the support reactions at A and B .
Solution: (1) We divide the area under the load into a rectangle and a triangle, areas of which we know the centroid.
Equilibrium requires

1
= 0 = 10 2.00 10
+ 0.75)
10 1.50 (1.50
+ 1.00)
+ B cos 30D 3.00 + B sin 30D 4.00
1.50


(1.50



2
15


centriod triangle
centroid rectangle
7.5

= 0 = Ay 10 + 15.77 sin 30D

72.50
= 15.77 k
4.598
Ay = 2.12 k

= 0 = Ax 15 7.5 + 15.77 cos 30D

Ax = 8.84 k

B=

F
F

71

(5.40)

Problems
5.1
y

(a)

(b)

A2
4.00

1.00

y= x

y
y =k x 2

A1

x
3.00

1.00

(a) Calculate the x - and y - coordinates of the centroids of areas A1 and A2 .


(b) Calculate the x - and y - coordinates of the shaded area.
Solution:
(a) x 1c = 0.600, y1c = 0.375, x 2c = 0.300, y2c = 0.750
(b) x c = 1.50, yc = 1.60
y

5.2

y
(b)

(c)

10

50

50

10

(a)

20

r = 10
x

x
30

40

30

x
20

30

Calculate the x - and y - coordinates of the centroids of the shaded areas.


Solution: (a) x c = 29.27, yc = 15.73

(b) x c = 23.33, yc = 22.22

5.3

(c) x c = 24.64, yc = 17.10

45D

1.50 m

1.50 m

2.25 m

2.25 m

Calculate the angle such that the eccentric artwork hangs in the horizontal position as shown.
Solution: = 39.6D
72

5.4

e = 0.12L

e=?

(a)

(b)

(a) Bricks of length L and equal weight are placed on top of each other each having an offset of e = 0.12L to the brick
below. How many bricks can be placed in this fashion until the brick-tower becomes unstable.
(b) If the brick-tower is to be 41 bricks high, find the maximum allowable offset e for stability of the tower.
Solution: (a) n = 8

(b)

e
= 0.024
L

5.5
0.70

3.30

0.30

2.00

0.30

3.00

[m]

x
2.00

2.00

0.50

1.50

Locate the centroid of the box section shown, a cross-section often used in bridge construction.
Solution: xc = 6 m, yc = 2.23 m
5.6

0.5

r = 2.5

r = 2.5

0.5

4r
3

Find the x - and y - components of the centroid of the shaded area (left). The figure (right) helps you locate the centroid of a
semi-circle.
73

33

5.7

16

13

x
33

12

30

Find the x - and y - components of the centroid of the shaded area.


5.8

4 kN/m

(a)

2 kN/m

(b)

2 kN

3 kN
B

A
A
3.00

[m]

1.50

4.00

1.50

[m]

4.00

30 kN/m

2 k/ft

1k/ft

2.00

(d)
3 ft

(c)

10 kN/m
30D

2 k/ft
5m

4 ft

Calculate the support reactions of the beams for the given loading.
Solution:
(a) Ax = 0, Ay = 8.0 kN , M A = 21.0 kNm (ccw)

(b) Ax = 0, Ay = 3.92 kN , B = 11.08 kN

(c) Ax = 7.44 k , Ay = 6.00 k , B = 1.44 k

(d) Ax = 76.97 kN , Ay = 100 kN , B = 134.7 kN

5.9

c =25 kN/m 3

z
y

1.0

[m]

10 (H + 1)
[kN/m]

A
2.0

1.0

2.0

Calculate the maximum allowable height H to prevent overturning of the concrete retaining wall shown (specific weight c ).
Consider a strip of length 1m in the y - direction.
Solution: H max 6.6 m

74

5.10

30 kN/m

15 kN/m
30D

3.00 m

3.00 m

2.00 m

3.00 m

Calculate the reactions at supports A and B for the given loading.


Solution: Ax = 24.25 kN , Ay = 31.56 kN , B = 41.5 kN 3
5.11

4 k/ft

3 ft

4 k/ft

2 k/ft

3 ft

2 k/ft

A
4 ft

4 ft

Calculate the reactions at support A for the given loading.


Solution: Ax = 18 k

Ay = 0

M A = 137.3 k-ft (ccw)

5.12
(a)

10 kN/m

20 kN
A

1.00 m

3.00 m

1.00 m

2.00 m

2.00 m

1.00 m

20 kN

3.00 m

1.00 m

10 kN/m

(b)
B

A
1.00 m

3.00 m

2.00 m

2.00 m

1.00 m

16 kN/m

(c) A

B
4.00 m

C
1.00 m

4.00 m

Find the reaction forces for the beams with internal pin(s).

75

3.00 m

1.00 m

Solution:

(a) Ay = By = C y = Dy = 20 kN

Ax = 0

(b) Ay = Dy = 18.33 kN

By = C y = 51.67 kN

Ax = 0

(c) Ay = 37.33 kN

By = 26.67 kN

Ax = 0

M A = 128 kNm (ccw)

25 k

5.13

2 k/ft
3 k/ft
A

(a)

(b)

10 ft

10 ft

15 ft

15 ft

20 ft

20 ft

3 k/ft
2 k/ft

(c)

2 k/ft
B

(d)

10 ft

15 ft

15 ft

20 ft

20 ft

Find the forces in columns 1, 2, 3 for the given loading. Note that the forces in the diagonal members are zero.
Solution:
(a) F1 = 25.0 k
(b) F1 = 37.5
(c) F1 = 60.0 k
(d) F1 = 25.0 k

= 25.0 k
= 7.5 k
= 10.0 k
=0

F2
F2
F2
F2

F3
F3
F3
F3

= 25.0 k
= 30.0 k
= 50.0 k
= 25.0 k

=compression, + = tension

5.14
z

3 k/ft

3 k/ft

3
1

3
2

5 ft

5 ft

Find the forces in columns 1, 2, 3 (spaced at 120 degrees) for the given loading. Think.

76

10 ft

5.15

5 kN/m

5 kN/m

10 m

5m

3m
A

4m

Line AB is pinned to the ground at A , supported by two cables and loaded as shown. Find the tension forces in the two
cables.
Solution: T1 = 46.58 kN, T2 = 64.75 kN

5.16
W

y
2

3
8 ft

6 ft

6 ft

4r
3

centroid of semi-circular area

6 ft

2 ft

8 ft

Three columns support a roof of weightW . (a) Locate the center of gravity of the roof with respect to the given
x -y coordinate system. The centroid of a semi-circular area is located as shown; (b) Find the resulting forces in the
columns.
Solution: (a) xC = 9.31 ft, yC = 10.00 ft, (b)F1 = 0.418W , F2 = 0.415W , F3 = 0.166W

77

5.17

Wtable = 300 lbs


B

a /2
a

a /2

Mrs./Mr. Tooheavy who presently weighs 200 pounds, wants to step on a table on three legs with square table top of weight
300 pounds. The legs are unable to resist any tension (uplift) forces and their weight is negligible. How many pounds does
Mrs./Mr. Tooheavy have to lose before the structure is "safe" against overturning regardless of her/his position on the table?
5.18
2 k/ft

6 ft

6 ft

10 k

B
10 ft

10 ft

Find the reaction forces at A and B of the structure for the given loading.
5.19
2 k/ft

3 k/ft

z
y

3
B
15 ft

5 ft

15 ft

Find the forces in links 1,2, 3 for the given loading and clearly indicate whether the links are in tension or compression. The
forces in the diagonal members are zero.

78

6. Trusses
6.1
6.1.1

Introduction
Historical development

from Encyclopaedia Britannica


"In engineering, a structural member usually fabricated from straight pieces of metal or timber to form a series of triangles
lying in a single plane. A truss gives a stable form capable of supporting considerable external load over a large span with
the component parts stressed primarily in axial tension or compression. The individual pieces intersect at truss joints, or panel points. The connected pieces forming the top and bottom of the truss are referred to respectively as the top and bottom
chords. The sloping and vertical pieces connecting the chords are collectively referred to as the web of the truss. Trusses
were probably first used in primitive lake dwellings during the early Bronze Age, about 2500 BC. The first trusses were
built of timber. The Greeks used trusses extensively in roofing, and trusses were used for various construction purposes in
the European Middle Ages. Andrea Palladio's I quattro libri dell'architettura (1570; Four Books on Architecture) contained
plans for timber trusses. A major impetus to truss design came in the development of covered bridges in the United States in
the early 19th century. Cast iron and wrought iron were succeeded by steel for railroad truss bridges. The two systems most
commonly used are the Pratt and the Warren; in the former, the sloping web members are parallel to each other, while, in
the latter, they alternate in direction of slope (see Fig. 6.4). Trusses are also used in many kinds of machinery, such as
cranes and lifts, and in aircraft wings and fuselages."
6.1.2
6.1.2.1

Load resisting behavior


General Remarks

Picture 6.1 Railroad and pedestrian truss bridges.


So far our equilibrium problems have been concerned with calculating reaction forces of a rigid body subject to externally
applied concentrated or distributed forces. All force were external to the body, a single free-body diagram was sufficient to
solve the problem. In this chapter we look at internal forces of truss structures, forces of action and reaction between connected members. We will look at several free-body diagrams, each for a different part of the structure. We wont learn any
new theory. The material covered in this chapter is a straightforward application of the concept of equilibrium developed in
previous chapters.
A structure composed of members connected at their ends is known as a truss. Simply speaking, we use trusses to span distances and safely transfer the forces applied over that distance to supports and from the supports into the ground or other
connecting structural elements. Bridges, roofs, cranes are common examples of trusses. Plane trusses are truss structures
whose members lie essentially in one plane. Picture 6.1 shows a conventional railroad steel truss bridge. Cross beams transfer the forces due to the weight of the train to the two trusses placed on each side of the bridge. The forces travel through the
truss into bearings placed on top of the abutments and from the abutments into the ground. Truss structures whose members
are not in a single plane are called space trusses (see Picture 6.3).

79

The truss members are commonly slender with little resistance to lateral load. All forces are applied to the connection (the
joints) and not to the members themselves. A joint is a connection of two or more members, which is usually assumed a
frictionless pin connection although the actual connection is welded, riveted or bolted (see Picture 6.2). Consequently, the
members connecting to the joints are allowed to rotate against each other and cannot transfer a moment. Consequently, the
force in the member is directed along the member and is either a tension force or a compression force.

Picture 6.2 Riveted and bolted connection of truss members.

Picture 6.3 Simple space truss.

Picture 6.4 Space trusses used in modern airport terminal construction.


The basis unit of a truss structure is a triangle, that is three members are joined at their ends to form a stable configuration.
Usually, the forces applied to the truss at the joints are large compared to the weight of the truss members such that we can
80

often neglect the self-weight. If the weight of the members is not negligible, we consider half the weight as external force at
each end joint. A triangle is rigid in the sense that it is non-collapsible. It can only move as a rigid body, i.e. all members
must synchronize as they move. A quadrilateral unit is collapsible, in the sense that even though we fix part of the unit, the
rest can still move (see Fig. 6.1).
does not change shape

changes shape

Fig. 6.1 Triangular unit (rigid, non-collapsible), quadrilateral unit (non-rigid, collapsible).

TENSION (T)

COMPRESSION (C)

or

Fig. 6.2 A tension or compression force acts on the member.


As illustrated in Fig. 6.2 the members of a truss are either in tension or in compression. On a free-body diagram of the
member, an arrow away from the member indicates that the member is in tension and an arrow towards the member
indicates that the member is in compression.
6.1.2.2

Trusses as a link in the structures load path

In a structure, loads travel along load paths from their point of application in the structure to the foundation and into the
ground. Trusses are often an essential component of the load path. Fig. 6.3 shows a simple structure in which the roof loads
(dead and occupancy loads) are transferred from the decking to the joists, from the joists to the trusses, from the trusses to
the walls, and finally from the walls into the foundation. The surface loads acting on the roof decking (indicated by the
many red arrows in Fig. 6.3(a) and (b) are resisted by line loads along the joists (the blue arrows in Fig. 6.3(b)). The line
loads of the joists are in turn resisted by point loads acting on the trusses (see Fig. 6.3(c)). The walls then support the point
loads acting on the trusses. Finally the wall loads are transferred into the foundation (not shown). We see that the reaction
forces acting on a member become the acting forces on other structural elements. For example, the blue line loads in Fig.
6.3(b) are the reaction forces acting on the decking. The red line loads in Fig. 6.3(c) are the same forces acting on the joists.
Likewise, the blue point loads in Fig. 6.3(c) are the reaction forces acting on the joists. The red point loads in Fig. 6.3(d) are
the same forces acting on the trusses. We establish the magnitudes of all these forces by applying the equations of equilibrium to several free-body diagrams. Regarding our discussion of trusses in this class, we assume that the load path analysis
has already been completed and the forces acting on the trusses are known. An in depth discussion of load paths and load
tracing is the subject of the Structural Systems (ARCE 226 and 371) classes.
Fig. 6.4 gives examples of common trusses. Fig. 6.5 shows two examples of space trusses.

81

(a)

(b)

(c)

(d)

Fig. 6.3 Basic load flow through different structural elements.

82

Warren truss

Pratt truss

Howe truss

K-truss

Bowstring truss

Pratt (roof) truss

Howe (roof) truss

Fig. 6.4 Several examples of plane trusses.

Fig. 6.5 Examples of space trusses.

83

6.1.2.3

Tension and compression forces and their relative magnitudes in simple trusses

(a) simply supported truss

(b) cantilever truss

max Tension

max Compression

Fig. 6.6 Structural model, tension and compression members and deflected shape in simply supported and cantilever trusses.
Fig. 6.6(a) shows a structural model of a plane truss similar to the two plane trusses located on each side of the bridge in
Picture 6.1. The truss is called simply supported, a pin and a roller supports are placed at the two ends. In the railroad bridge
in Picture 6.1, pin and roller supports (see Picture 3.3) represent the bearings placed on the abutments. The main structural
elements of the truss are the bottom chord, the top chord and the diagonals. Vertical forces are applied to the joints of the
top chord. Also shown in Fig. 6.6(a) is the deflected shape of the truss along with a graphical representation of the member
forces. We observe that the member forces vary throughout the structure. The top chord shortens and is thus in compression,
i.e. all forces in the top members are compression forces. The bottom chord elongates and is thus in tension. The diagonal
members are alternately tension and compression. Compression and tension forces in the longitudinal members are largest
at mid-span whereas the forces in the diagonal members take on their maximum at the supports.
Fig. 6.6(b) shows a cantilever truss. A cantilever truss has a free end (no supports), while the other end is supported both
against translation and rotation. For this type of truss the bottom chord shortens and is in compression and the top chord
elongates and is in tension. Tension and compression forces in both longitudinal and diagonal members are largest at the
support. Note that where the diagonals change direction, the sense of the member forces changes.

84

6.2

Statical determinacy, indeterminacy and instability of trusses

As already discussed in Chapter 3 for simple rigid bodies, we need to identify whether a truss is statically determinate,
statically indeterminate, or unstable. Fig. 6.7-Fig. 6.9 depict examples for each of those properties.
The truss in Fig. 6.7 is constructed by adding eight triangular panels. All members and all supports are necessary for stability of the truss, i.e. if we remove one member from the structure it becomes unstable and collapses under applied loads. The
truss is statically determinate. We will learn how to analyze this type of trusses in subsequent sections of this chapter.
The trusses in Fig. 6.8 are statically indeterminate. The structure in Fig. 6.8(a) has one more support than necessary for stability. Since the indeterminacy is with respect to the reactions, the truss is termed statically indeterminate externally. The
structure in Fig. 6.8(b) has more members than necessary for stability. We can remove one of the top three members without
sacrificing the stability of the structure. Since in this case the indeterminacy is with regard to the numbers of members, the
truss is called statically indeterminate internally. Analysis of indeterminate structures is beyond the scope of this class.
The trusses in Fig. 6.9 are unstable. The structure in Fig. 6.9(a) is externally unstable because it has no support that resists
horizontal movement (we would have to turn one roller support into a pin support to make the truss statically determinate).
The structure in Fig. 6.9 (b) is internally unstable because it has too few members. More precisely, the square panel is unstable since it has no diagonal (we would have to add one diagonal to make the truss statically determinate). Both structures
undergo excessive displacements without forces being applied to it, a phenomenon already illustrated in Fig. 6.1 for a single
quadrilateral unit. Figure Fig. 6.9 illustrates those instability modes.
Although we will discuss formulas to calculate the degree of indeterminacy and instability in subsequent classes, the best
method by far to assess those properties is to use intuition, inspection, insight and experience.

Fig. 6.7 Statically determinate truss.

(a)

(b)

Fig. 6.8 Statically indeterminate trusses.

(a)

(b)

Fig. 6.9 Unstable trusses with collapse modes.


85

6.3
6.3.1

Methods of joints
Procedure

The method of joints is a method to determine members force of truss structures by considering equilibrium of individual
joints. Since we assume the line of action of all forces applied to the joint to pass through the joint, we deal with concurrent
forces. Correspondingly, we have two equilibrium equations since we satisfy moment equilibrium automatically. We start
the analysis at a joint where at least one known force and not more than two unknown forces exist. We draw a free-body
diagram of the joint and designate the forces by letters defining the ends of the member. If we draw the force arrows on
the same side of the pin as the member, tension is indicated by arrow away from the joint, compression indicated by
arrow towards the pin. For each joint, we solve two equilibrium equations, Fx = 0 and Fy = 0 . Again, we cannot
use a moment equation since all forces involved are concurrent at the joint. A negative answer for an unknown force
indicates that its action is in the sense opposite to that assigned in the free-body diagram. This happens frequently, when the
correct physical sense is not known initially, so that an arbitrary assignment becomes necessary. We then progress from
joint to joint always selecting a joint with no more than two unknown forces.

F
C
(a)

(b)

Fig. 6.10 Simple truss: (a) mechanical model, (b) free-body diagram of entire structure.
When analyzing the truss in Fig. 6.10 we start at joint C , where the applied force F is known and member forces CA and
CB are unknown. We solve the equilibrium equations for those two forces. We then proceed to node B . At node B , we
know force BC (= CB ) from the previous step and member force BA and the vertical external reaction force are the unknowns. We solve the equilibrium equations at joint B and finally move to joint A to find the external reaction forces at
pin A . As a check, we can use the free-body diagram in Fig. 6.10(b) and calculate the reaction forces by applying the equilibrium equations to the whole truss.
Fig. 6.11 presents an exploded view to illustrate the concepts of tension and compression in the members, internal and external forces, and joints and members.

applied force
joint

compression member

joint

compression member

tension member
joint

reaction force

reaction force

Fig. 6.11 External and internal forces for simple truss.

86

6.3.2

Example 6.1
30D
4000 N
B

3.00

C
500 N

4.00

4.00

Fig. 6.12 Example truss.


Problem: Use the method to joints to find the member forces and the support reactions of the truss structure in Fig. 6.12.
Solution: (1) We start the analysis by looking at equilibrium of joint C . Three forces act at joint C , the externally applied
force, the yet unknown member force CB and the unknown member force CD . Fig. 6.12 shows the corresponding free-body
diagram for this joint. It is good practice to assume all unknown member force to be in tension such that they act away from
the joint in the free-body diagram. Note that joint C is the only joint with only two unknown forces. As before, known and
unknown forces are shown in red and blue color, respectively.
30D

2965

4000 N
B

3286

CB

CD

3.00

1286

C
500

sin =

3
= 0.6
5

cos =

4
= 0.8
5

C
500 N

4.00

4.00

Fig. 6.13 Free-body diagram for joint C .


Equilibrium in the vertical direction gives

= 0 = 500 + CB sin CB =

500
500
=
= 833.3
sin
0.6

(6.1)

Equilibrium in the horizontal direction yields

= 0 = CB cos CD CD = (833.333) 0.8 = 666.7

(2) Next, we can look at joint D . Since we have calculated force CD , joint D has two unknown forces, the force in
members DE and DB (see Fig. 6.14).

87

(6.2)

30D
4000

DB
DE D
666.7
C

D
E

500

Fig. 6.14 Free-body diagram for joint D .


Equilibrium at joint D gives

= 0 DB = 0

= 0 = 666.7 DE DE = 666.7

(6.3)

(3) The next step is to look at joint B . All forces acting on joint B are known from the analysis of joints C and D except
member forces BA and BE . Note that since the results for force BC was negative, the 833.3-force is a compression force
and hence points toward the pin. Recall that we never write negative numbers on the free-body diagram. The free-body
diagram for joint B and the corresponding equilibrium equations are shown below in Fig. 6.15
30D
4000

30D

4000
BA B

BE

D
E

833.3

500

Fig. 6.15 Free-body diagram for joint B .

= 0 = 4000 cos 30D + 833 sin BE sin

BE =

1
(3464 + 833.3 0.6) = 4941
0.6

(6.4)

= 0 = BE cos BA 4000 sin 30 833.3 cos BA = 2000 (4941) 0.8 833.3 0.8 = 1286

(4) We now proceed to joint E . Since we have calculated forces EB and ED before, the unknown forces at joint E are member force EA and the support reaction E x . Note that equilibrium at joint B gave us a negative value for BE such that BE is a
compression force pointing towards joint E .
30D
4000

EA

Ex
D
E

500

Fig. 6.16 Free-body diagram for joint E .

88

4941

666.7

= 0 = E x + 666.7 4941 0.8

E x = 3286

= 0 = 4941 0.6 + EA

EA = 2965

(6.5)

We have now calculated all member forces of the truss as well as the support reaction E x . To determine the two remaining
support reactions Ax and Ay , we look at joint A .
30D
4000

Ay

Ax

1286
2965

D
E

500

Fig. 6.17 Free-body diagram for joint A .

= 0 = Ax + 1286 Ax = 1286

= 0 = Ay + 2965 Ay = 2965

(6.6)

We calculated the three support forces Ax , Ay , and E x by applying the equilibrium equations to joint A and E . In analyzing
the truss, we havent used the equilibrium equations for the whole truss. Doing so provides a check of the results. Applying
the equilibrium equations to the free-body diagram of the entire truss in Fig. 6.18 yields

F = 0?
F = 0?
M = 0?
x

3286 1286 4000 sin 30D = 0

500 4000 cos 30D + 2965 = 1 0

ok

500 8 4000 cos 30 4 + 4000 sin 30 3 + 1286 3 = 2 0

30

2965
1286

ok
ok

4000 N
B

3.00

3286

500 N
4.00

4.00

Fig. 6.18 Free-body diagram of entire truss for equilibrium check (all forces known).
Finally, we summarize the results by drawing the truss with the forces written next to the corresponding member.
30D
1286

2965 (T)

3286

2965

4000

1286 (T)

4941 (C)

833(C)

0
666.7 (T)

666.7 (T)
500

Fig. 6.19 Summary of member forces (forces in N).


89

(6.7)

6.4

Special loading conditions

6.4.1

Discussion

We should devote special attention to two special loading conditions frequently encountered in truss structures.
F3

F1
F2

F2

x
x

90D

90D

90D
(b)

(a)
F1

Fig. 6.20 Special loading conditions


In Fig. 6.20(a), three truss members are connected to a joint. Two members are collinear. From equilibrium in the direction y perpendicular to the collinear members we conclude that the force in the third member is zero (see Eq. 6.8). With
member 3 a zero-force member, equilibrium in the x -direction requires that the forces in members 1 and 2 are equal (see
Eq. 6.8).

= 0 = F3 cos F3 = 0

= 0 = F1 F2 F1 = F2

(for any )

(6.8)

In Fig. 6.20(b), two truss members in different directions are connected to a joint. From equilibrium in the direction x perpendicular to member 2 we conclude that the force in member 1 is zero. Likewise, equilibrium in the direction x
perpendicular to member 1 gives us zero force in member 2.

= 0 F1 = 0

= 0 F2 = 0

(6.9)

Note that the forgoing is true only if there is no external load applied to the joint.
Picture 6.5 shows a balcony structure to illustrates the phenomenon in Fig. 6.20(b). Members 1 and 2 connect at a joint
where no external force is applied. The force in both member 1 and member 2 is zero and all the vertical force is supported
in the diagonal. Strictly speaking, there are always small forces in the members due to self-weight of the members.

Member1
Member 2

Picture 6.5 Timber balcony structure with zero force members.

90

6.4.2

Example 6.2
J

I
H

K
N

G
A

Fig. 6.21 Example 6.2.


Problem: Identify all zero-force members of the truss in Fig. 6.21.
Solution: Using the information in Fig. 6.20, we always solve problems like this by inspection. We observe that joints
I , K , and M represent the condition discussed in Fig. 6.20(a), i.e. one member connects to two collinear members. We thus
conclude that member forces IN , KO, and EM are zero. With those zero-force members, however, joints N ,O, and E also
become special loading conditions such that forces HN , LO, and EL are zero. Finally, with force OL and EL equal to zero,
force DL is also zero since joint L has become a special loading condition. Fig. 6.22 summarizes the results by highlighting
the zero-force members.
J

I
H

K
N

G
A

Fig. 6.22 Example 6.2, zero-force members.

6.5

Methods of sections

6.5.1
Procedure
In the method of joints we used two of the three equilibrium equations. We couldnt make use of the moment equation.
since all forces were concurrent at the joint. The method of sections, developed by the German engineer A. Ritter in 1862,
makes use of moment equation by selecting an entire section of the truss for the free body in equilibrium under action of
non-concurrent forces. The big advantage of the method of sections is that we can find the force in almost any member
directly from an analysis of a section that cuts through that member. Hence, we dont need to move from joint to joint until
we reach the member in question. Usually, we should not cut through more than three members, since there are only three
independent equilibrium equations. Note that the forces of the intact members are not involved in the analysis. To
illustrate the method of sections we use the truss discussed before to discuss the method of joints.

91

6.5.2

Example 6.3
30D
A

AB

AB

4000 N

3.00

3.00

BE
BE

DE

DE

C
4.00

500 N

Fig. 6.23 Two sections of truss structure.


Problem: Use the method of sections to find member forces AB, BE and DE .
Solution: We cut the truss structure into two parts at an appropriate location. The two portions of the truss must be in equilibrium. Hence we have to apply to each cut member the force exerted on it. As in the method of joints, these forces, either
tension or compression, will always be in the direction of the member. Since we have only three independent equilibrium
equations, we must not cut through more than three truss members. Fig. 6.23 shows the two sections of the truss structure
generated by the cut. The left-hand section is in equilibrium under the action of the three support forces Ax , Ay , and E x and
the member forces AB, BE and DE . The right-hand section is in equilibrium under the action of the two externally applied
forces and the member forces AB, BE and DE . We can use either section to calculate the member forces. Since we dont
know the support forces at A and E it is advantageous to look at the right-hand section and write the equilibrium equations.
We observe that all three forces have components in the horizontal direction, such that it is not a good idea to start with
equilibrium in the horizontal direction. It is clever to formulate vertical equilibrium first since member force BE is the only
unknown force contributing to vertical equilibrium. We obtain

= 0 = BE sin 4000 cos 30D + 500 BE =

500 3464
= 4940
0.6

(6.10)

Next, we find member force DE . Recognizing that both AB and BE pass through B it is elegant to select joint B as the reference point for a moment equation which gives

= 0 = DE 3 + 500 4

DE =

2000
= 666.7
3

(6.11)

Finally, we find member force AB . Recognizing that both DE and BE pass through E it is advantageous to select joint E as
the reference point. Note that although point E lies on the left-hand section of the structure, we still apply the equilibrium
equations to the right-hand side. Remember that the reference point for the moment equation does not have to lie on the
body to which we apply the equilibrium equations. We should exercise great caution to include all forces that have a
moment about E .

= 0 = 4000 cos 30D 4 + 4000 sin 30D 3 + 500 8 + AB 3 AB =

3856
= 1285
3

(6.12)

Since we have assumed BE to be in tension and the numerical result is negative, we conclude that BE is a compression
force. Note that we have calculated all three unknown forces independently of each other (three equations with one unknown each), such that an error in one equation does not influence other equations. Also note that we have not used force
equilibrium in the horizontal direction to calculate the unknowns. This may now serve as a check

= 0 ? AB DE BE cos 4000 sin 30D = 1286 666.7 2000 + 4940 0.8 = 0

92

ok

(6.13)

6.6

Summary

In the method of joints, we apply the two force equilibrium equations to each node. Since the member forces are concurrent at the joint, we can not make use of a moment equation. Usually, we prefer this method if we are interested in all
member forces. Member forces determined at a joint are transferred to adjacent joints and treated as known quantities.
In the method of sections, we cut through several members and consider an entire portion of the truss as the body in
equilibrium under non-concurrent forces. If we are only interested in a few member forces, this method is typically more
efficient. In many cases, moment equations are particularly advantageous. We should select a reference point through which
as many unknown forces as possible pass. As in the method of joints, it is probably best for us to assume all forces in
tension and let the algebraic sign of the results determine the actual direction of the force.

6.7

Example 6.4
20 kN

15 kN

2.00

2.00

10 kN
C

[m]
3.00

B
2.00

4.00

4.00

2.00

Fig. 6.24 Example 6.4.


Problem: Use the method of sections to find member forces AC , AD,CE,CG and FG .
Solution: In this example, we use the method of joints to find forces AC and AD and the method of sections to find forces
CE ,CG and FG .
(1) We start with calculating the reactions by looking at a free-body diagram of the whole structure.
20 kN

15 kN

2.00

2.00

10 kN

[m]
3.00

Ax

A
Ay

B
2.00

4.00

4.00

2.00

Fig. 6.25 Example 6.3, free-body diagram of whole structure.


93

Equilibrium requires

M = 0 = B 12 20 6 15 2 10 3
F = 0 = A + 15 35
F = 0 = 10 A

B = 15

Ay = 20

(6.14)

Ax = 10

(2) With known reaction forces we isolate joint A (see Fig. 6.26) to find members forces AC and AD
AD

AC
56.31D

10

20

Fig. 6.26 Free-body diagram of joint A


3
= tan1 = 56.31D
2

F
F

sin = 0.8321

cos = 0.5547

= 0 = 10 + AC 0.5547

AC = 18.03

= 0 = 20 + 18.03 0.8321 + AD

AD = 35.00

(6.15)

(3) Finally, we cut through the three members CE,CG and FG whose forces are desired and apply the equilibrium equations
to one of the two resulting free-bodies (see Fig. 6.27).
2
= tan1 = 26.57 D sin = 0.4472 cos = 0.8944
= 45D
4
MC ,I = 0 = FG 0.8944 2.00 10 3.00 20 2.00 FG = 39.13

MG,II = 0 = EC 0.8944 2.00 + 15 6.00

CE = 50.31

CG = 14.14

= 0 = 39.13 0.8944 GC 0.7071 50.31 0.8944

(6.16)

20 kN

10 kN
C

I
CG

II

CE

CE

[m]
3.00

CG

2.00

FG

2.00

FG
15 kN

10 kN

20 kN

B
2.00

4.00

4.00

2.00

15 kN

Fig. 6.27 Free-body diagrams of cut sections


(4) We summarize the results
AC = 18.0(T), AD = 35.0(C),CE = 50.3(T),CG = 14.1(C), FG = 39.1(C) [kN]

(6.17)

Note: We can also find force CE by looking at free-body diagram I and FG by looking at free-body diagram II. Can you
write the equations and get the results of Eq. (6.16)?
94

6.8
6.8.1

The 12-node structure


Building the model

Fig. 6.28 12 Columns on 4x3 Grid; Columns connected by 17 horizontal members


As we have learned in Chapter 1 of the class, a point in space has 3 motions. The 12 columns stop the 12 vertical motions.
Next we build 17 horizontal members connecting the top of the columns. We have now prevented 17+12=29 out of the 36
possible motions. Fig. 6.29 depicts the seven motions remaining

Fig. 6.29 Remaining motions after building columns and horizontal members

95

We can stabilize the structure by adding seven diagonal members, one in each wall plane.

Fig. 6.30 Stable structure with 36 members (12 columns, 17 horizontals, 7 wall diagonals)
For architectural reasons, we often want diagonal-free interior space. We thus move the four interior diagonals into the roof
plane.

z
y

Fig. 6.31 Stable structure with 36 members (12 columns, 17 horizontals, 3 wall diagonals, 4 roof diagonals)
The roof surface now acts like a fixed plane, i.e. that is the points forming the plane are fixed relative to each other. The
three in-plane motions of the roof plane relative to the ground (translations in the x -and y -directions and rotation about
the z -axis) are prevented by the three wall diagonals. The next section provides illustration how four diagonal members
stabilize a 2x3 panel plane for in plane motion.

96

6.8.2

Stabilizing the roof plane

Fig. 6.32 Roof plane without diagonals (unstable, four independent motions possible)

Fig. 6.33 Roof plane with one diagonal (unstable, three independent motions possible)

97

Fig. 6.34 Roof plane with two diagonals (unstable, two independent motions possible)

Fig. 6.35 Roof plane with three/four diagonals (unstable/stable).


It is important to note that the arrangement of diagonals is not arbitrary. Fig. 6.36 shows three arrangements of four
diagonals, that do not prevent all motions.

WRONG
WAY

Fig. 6.36 Roof plane with four diagonals (unstable, one motion possible).

98

Fig. 6.37 shows all possible combinations to stabilize a plane of 2x3=6 panels with four diagonals

RIGHT
WAY

Fig. 6.37 Alternatives for stabilizing 2x3 roof plane with four diagonals.

99

6.8.3

Analyzing the roof plane (Example 6.5)

L
H

2F

A
B

Fig. 6.38 Plane truss model for roof plane (rollers represent wall diagonals).
Problem: For F =1k , find all member forces in the plane truss structure above. All angles are 30, 60 or 90 degrees.

2
1
A

Fig. 6.39 Free-body diagram of whole structure.


Solution:
To find the member forces, we use the method of joints. We start with finding the support reactions by applying the three
equilibrium equations to the structure as a whole

F = 0 = I 2 11
M = 0 = 4 2 tan 30
F = 0 = A + D

I =4

2 tan 30 2 1 tan 30 + D 3

D = 0.770

(6.18)

A = 0.770

Next, we find as many member forces as possible by inspection. Based on the discussion in Section 6.4 (special loading
conditions) we identify the zero-force members EI ,GK , and HL .

A
0.770

0.770

Fig. 6.40 Zero-force members by inspection.


Also by inspection, we find the forces in members KL, JK , CD, DH .
IJ = 4

KL = JK = CD = 1

DH = 0.770

100

(6.19)

L
H

A
0.770

0.770

Fig. 6.41 Summary of member forces found by inspection.


Since we have to find the forces in all members of the structure, we use the method of joints. Recall from the discussion of
the method of joints that we start the analysis at a joint where at least one known force and not more than two unknown
forces exist. We select joint H .
I

0
H

G
D

A
0.770

GH

0.769

CH

0.770

Fig. 6.42 Isolating joint H and corresponding free-body diagram.


Equilibrium requires

F
F

y
x

= 0 = 0.769 CH sin 30D


= 0 = 2 GH CH cos 30

CH = 1.539

(6.20)

GH = 3.333
L

A
0.770

0.770

Fig. 6.43 Known member forces after considering joint H .


Next, we consider joint C .
I

L
H

D
B

CG
1.539

BC

A
0.770

1
0.770

Fig. 6.44 Isolating joint C and corresponding free-body diagram.


101

1.000

Equilibrium requires

F
F

= 0 = BC 1.000 + 1.539 cos 30D

= 0 = CG + 1.539 sin 30

A
0.770

(6.21)

CG = 0.769

BC = 0.333

0.770

Fig. 6.45 Known member forces after considering joint C .


We move on to joint G .
I

FG

A
0.770

BG

3.333
0.769

0.770

Fig. 6.46 Isolating joint G and corresponding free-body diagram.


Equilibrium requires

F
F

= 0 = 0.769 BG sin 30D

= 0 = FG BG cos 30 3.333

BG = 1.539

0.770

A
B

FG = 4.666

0.770

Fig. 6.47 Known member forces after considering joint G .

102

(6.22)

Joint B
I

1
BF

1.539

A
B

0.770

AB

0.333

0.770

Fig. 6.48 Isolating joint B and corresponding free-body diagram.


Equilibrium requires

F
F

= 0 = AB + 0.333 + 1.539 cos 30D


= 0 = BF + 1.539 sin 30
I

(6.23)

BF = 0.769

A
0.770

AB = 1.666

0.770

Fig. 6.49 Known member forces after considering joint B .


Joint A
I

1
AE

A
0.770

AF

1.666
0.769

0.770

Fig. 6.50 Isolating joint A and corresponding free-body diagram.


Equilibrium requires

F
F

= 0 = 1.666 + AF cos 30D


D

= 0 = AE + AF sin 30 0.769

AF = 1.924
AE = 1.732

103

(6.24)

A
0.770

0.770

Fig. 6.51 Known member forces after considering joint A .


Joint E
I

EJ

D
B

EF

A
0.770

1.732

0.770

Fig. 6.52 Isolating joint E and corresponding free-body diagram.


Equilibrium requires

F
F

y
x

= 0 = 1.732 + EJ sin 30D


= 0 = EF + EJ cos 30

EJ = 3.464

A
0.770

(6.25)

EF = 3.000

0.770

Fig. 6.53 Known member forces after considering joint E .


Joint F

0.770

FJ
3.000 F

A
B

1.924

0.770

Fig. 6.54 Isolating joint F and corresponding free-body diagram.

104

4.666
0.769

Equilibrium requires

F
F

y
x

= 0 = FJ + 0.769 + 1.924 sin 30D

FJ = 1.732

= 0 = 3 + 1.924 cos 30 4.666

(6.26)

ok

Finally, we check equilibrium at joint J .

4.000
3.464

1.000
1.732

Fig. 6.55 Free-body diagram of joint J (equilibrium check).


Equilibrium requires

F
F

= 0 = 4.000 1.000 3.464 cos 30D


= 0 = 1.732 3.464 sin 30

ok

(6.27)

ok

We conclude the analysis by providing a summary of the member forces.

1.666 T

Compression

Tension

Fig. 6.56 Graphical and numerical summary of member forces and deflected shape.

105

39

1.5

0.333 T

0
3.333 C

39

1.5

1.000 C

0.769 C

24

1.9

4.666 C

0.769 C

1.732 T

3.000 C

1.000 C

0.769 C

64

3.4

1.000 C

1.732 C

4.000 C

6.8.4

Trusses for vertical loads (remove interior columns)

Fig. 6.57 Three systems of vertical trusses replace interior columns (secondary members shown on the right provide
stability for additional joint).

106

Problems
6.1

3.00

2k

[ft]
2.00

4k

2.00

2.00

2.00

Calculate the force in each member of the truss above.


Solution:
AB = 2.83(T)
6.2

Tension

AD = 1.50(C)

BD = 1.50(T)

BE = 3.30(T) CE = 3.30(C)

D
2.00

60 kN

BC = 1.83(T)

E
2.00

50 kN C

40 kN

F
2.00

A
3.00

Find the members forces AF , BE , BF , and CE of the truss.


Solution: AF = 61.3 (T) BE = 55.9 (T) BF = 65.0 (C) CE = 50.0 (C)
6.3

Find the zero-force members for the truss and the given loading.
107

Compression

DE = 3.67 (C)

10 kN A

B
2.00

6.4

20 kN

2.00

[m]

2.00

2.00

2.00

2.00

Determine the forces in members AB and AD . (Hint: Combine methods of joints and section) .
Solution:
AB = 40.0 (T)

AD = 70.7 (C)

6.5

For the truss shown find the member with the largest:
(a) diagonal compression force, (b) diagonal tension force, (c) longitudinal compression force, (d) longitudinal tension
force, and calculate the corresponding forces. All applied forces have 2 k magnitude. All truss panels are 6 ft by 6 ft.
Solution:
(a)28.28 (C)

(b)14.14 (T)

(c) 90.0 (C)

(d)110.0 (T)

6.6

Repeat problem 5.5 for the truss structure above. The applied forces have 2 k magnitude. All triangles are 6 ft high and 6 ft
wide.
Solution:
(a)10.06 (C)

(b)10.06 (T)

(c)25.0 (C)

(d)24.5 (T)

108

6.7
7 kN
C

90D

14 kN

90D

90D

90D
6m

4m

10 kN

4m

Find the member forces of the truss for the given loading.
Solution:

AB = 9.0(T)

BC = 9.0(T)

CD = 19.5(T)

DE = 19.5(T)

AF = 39.0(C)

EF = 25.0(C)

BF = 0

DF = 14.0(C)

2 kN

2 kN

2 kN

2 kN

1kN

1 kN

2 kN

2.00

2.00

6.8

CF = 27.8(C)

2.00

2.00

2.00

2.00

D
2.00

2.00

(a) Determine the forces in members HI , CH , and BC .


(b) Determine the forces in members KL , KN , DN , and DE .
Solution:
(a) HI =8.00 kN (C), CH =4.24 kN (T), BC =5.00 kN (T)
(b) KL = 2.50 kN (C), KN = 2.12 kN (C), DN = 2.12 kN (T), DE = 2.50 kN (T)

109

6.9
F

G
2 ft

I
2 ft

C
5k

5k
4 ft

2 ft

2 ft

4 ft

Determine the forces in members EF , EH , AH and AB .


Solution:
EF =5.00 kN (C), EH =5.38 kN (T), AH =8.08 kN (C), AB =7.50 kN (T)

Tension

6.10

Compression

120 kN
C

30 kN

The truss shown is composed of 45 degrees right triangles. (a) Find the forces in members AB and AC ; (b) Find the forces
in members DF and DE .
Solution: AB = 42.42 (T ), AC = 100 (C), DE = 150 (T), DF = 70.71(T)

110

6.11

30 k

E
12 ft

10 k B

G
12 ft

12 ft

20 k C

H
16 ft

16 ft

For the truss shown: (a) determine all zero-force members; (b) calculate the forces in members FG and GJ
Solution: FG = 11.25 k (C), GJ = 31.25 k (C)
6.12

F
F
F
F

We model the roof plane of the above 3-D truss-structure by a plane truss model whose three roller supports represent the
three wall diagonals in the three-dimensional model. Identify the zero-force members in the plane truss and find the forces
in the non zero-force members. Angles are either 45 or 90 degrees.
Solution:
Member Forces (multiples of F )

1.00 T

41

1.

0
2.00 C

41

1.

1.00 C

41

1.

3.00 C

1.00 C

2.00 T

2.00 C

1.00 C

1.00 C

82

2.

1.00 C

2.00 C

3.00 C

1.00 C

111

P /2

6.13

P
P
P /2

P /2
P

P
P /2

Below are 100 different alternatives to stabilize a 4x3=12 panel- (5x4=20 node-) plane with the minimum number of
diagonals. All cases have the same loading and support conditions as the structure above. For P=20 k , analyze the combination that corresponds to the last two digits of your social security number. If your social security number ends on 00,
work on combination 100. All angles are 30, 60 or 90 degrees. Use inspection to identify as many member forces as
possible. Find all member forces. Clearly indicate whether a member is in tension or compression.
1

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

112

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

113

6.14

10 k

I 4 B

z
y

5
H

O, D

A 1

8 ft

Six links support a truss with six members in form of a tethrahedron. Points A, B,C form an equilateral triangle in the
x -y plane. The projection O of point D onto the x -y plane is the center of a circle of radius 8 ft through A, B,C . The height
of the truss (distance OD ) is 12 ft. Find the forces in the six links as well as the forces in the six truss members. Clearly
indicate whether the forces are tension or compression.
Solution:
Forces in links
1
2
0
0

3
3.33

4
0

5
3.33

6
3.33

Forces in truss members


1
2
3
4
5
6
4.01
4.01
4.01
1.28
1.28
1.28
Note: positive = tension, negative = compression, forces in kips
6.15
Repeat problem 5.12 for a 10 k force applied at point D in the x -direction. The three diagonal links are at an angle of 45
degrees.
Solution:
Forces in links
1
2
14.14
0

3
0

4
0

5
5.00

6
5.00

Forces in truss members


1
2
3
4
5
6
12.02 6.01 6.01 1.93 1.93 1.93
Note: positive = tension, negative = compression, forces in kips

114

6.16

F4
7

F7

F6

12

16
4
1

15
3

F5

F3

6
18 11 14
2
5
17 13

F1

3 ft
10 F2
3 ft
5

3 2

6 ft

9 ft
4 ft

4 ft

Six links support a three-dimensional truss structure. The truss consists of 18 members: Members 1-8 are the horizontal
members at the bottom and top, members 9-12 are the four verticals, members 13-16 are the four wall diagonals and members 17-18 are floor and roof diagonals. Find the forces in the six links as well as the forces in the 18 truss members. Clearly
indicate whether the forces are tension or compression. Consider the forces separately. Each force has 10 k magnitude.
Solution:
Forces in links
Link F1
F2
F3
F4
F5
F6
F7
1
12.5
0 12.50
0
0 12.5
0
2
0
0 8.33
0
0 8.33 12.5
3
4.17 5.00 9.17 5.00 10.0 9.17 2.50
4
0 12.5 8.33 12.5
0 8.33
0
5
3.33 7.50 1.67 7.50 10.0 1.67
0
6
0 5.00
0 5.00 10.0
0 5.00
Forces in truss members
Truss member
F1
F2
F3
F4
F5
F6
F7
1
0
0
0
0
0
0
0
2
0 5.00 3.33 5.00 10.00 3.33
0
3
0
0
0
0
0
0
0
4
0 5.00 3.33 5.00 10.00 3.33 10.00
5
0 7.50 5.00 7.50 15.00 5.00
0
6
0 10.00
0
0
0
0
0
7
0 7.50 5.00 7.50 15.00 5.00
0
8
0
0
0
0
0
0 10.00
9
0 2.50 1.67 2.50
5.00 1.67 5.00
10
3.33 2.50 1.67 2.50 5.00 1.67
0
11
0 2.50 1.67 2.50 5.00 1.67
0
12
0 2.50 1.67 2.50 5.00 1.67
0
13
10.54 7.91 5.27 7.91 15.81 5.27
0
14
0 5.59 3.73 5.59 11.18 3.73
0
15
0 7.91 5.27 7.91 15.81 5.27
0
16
0 5.59 3.73 5.59 11.18 3.73 11.18
17
0 9.01 6.01 9.01 18.03 6.01
0
18
0 9.01 6.01 9.01 18.03 6.01
0
Note: positive = tension, negative = compression, forces in kips
115

6.17

max Tension

max Compression

For the 12-node structure, three different vertical trusses replace the two interior columns. Two forces of 10 k magnitude are
applied as shown. For the three truss designs shown, find the forces in all members of the structure. The structure has a
height of 15 feet; each structure panel has a length of 18 feet in the two horizontal directions. The depth of the vertical trusses is 3.5 feet. Qualitative results provided in the figure should help you solve the problem.

116

7. Stress, strain, Hookes law and axially loaded structural members


7.1

Axial stress

7.1.1
Discussion
So far we have been concerned with calculating forces in links that connect a structure to the ground (external forces) and
forces in links (members) that connect different parts of a structure (internal forces). We have used the equations of
equilibrium to calculate these forces. Since force is not a suitable measure for the demand on a structural member, we will
in this chapter look at the intensity of a force, i.e. the force per unit area. Whether or not a member fails under a given
loading depends on whether the material is able to withstand the distributed forces acting on the cross section.
Throughout this chapter, we assume that the member force, also referred to as axial or normal force is applied along the
centroid of the cross section. In that case (see Fig. 7.1), the distribution of forces over the cross section is uniform. Considering that the sum of these forces must equal N and denoting the force per unit area by the Greek letter (sigma), we
have
=

N
A

(7.1)

We call this force per unit area the axial or normal stress. Axial stress is either a tensile or compressive stress. The common
sign convention is that a positive stress values indicates a tensile stress (the member is intension) and a negative sign indicates a compressive stress (member in compression). In the SI system of units, we measure stress in N/m2 . 1 N/m2 is
called 1 Pascal. Since one Pascal is a small pressure, other measures like kPa ( 103 Pa ), MPa ( 106 Pa ) or GPa ( 109 Pa ) are
more common. In the US system, stress is measured in pounds per square inch (psi), or kips per square inch (ksi). Familiar
examples of such a uniform distribution of forces are tire pressure (30-50 psi) or hydrostatic pressure (e.g. 13 psi at 30 ft
water depth).
P

stress
concentration

uniform
stress distribution

N
=

N
A

stress
concentration

uniform
stress distribution

Tension

Compression

Fig. 7.1 Illustration of axial force P , uniform axial stress and stress concentrations.
Equation (7.1) is only valid only if the axial stress is uniformly distributed over the cross-sectional area of the member. This
requirement is satisfied if the axial force P acts through the centroid of A and the location of interest is sufficiently far away
from areas of stress concentration (see Fig. 7.1). Stress concentrations develop if the axial force P is applied over an area
smaller than the cross-sectional area or sudden changes in cross-sectional area exist.
We can apply the concept of stress to perform elementary design of tension and compression members. Knowing the axial
force N to be transmitted and the allowable axial stress allow of the material, we obtain the required cross-sectional area
Areq from the relation
117

Areq =

N
allow

(7.2)

In order to analyze a stress problem we first have to use the equations of equilibrium to find the force in that member. We
then use the cross-sectional area of the member to calculate the stress according to Eq. (7.1). We discuss this strategy by an
example.
7.1.2

Examples

Example 7.1
200 kN

100 kN
3.00 m

A
4.00 m

Fig. 7.2 Example 7.1.


Problem: Calculate the axial stress in each member of the truss in Fig. 7.2. Tension members have a cross-sectional area
A=30 cm2 . Compression members have cross-sectional area A=50 cm2 .
Solution:
1. Statics: We first need to calculate the member forces. We use the method of joints as discussed in Chapter 6 of this class.
200 kN

BC

AC

100 kN

Fig. 7.3 Free-body diagram to find member forces


Applying the equations of equilibrium to the free-body diagram joint C (see Fig. 7.3) gives

= 0 = 200 AC sin

Fx = 0 = 100 AC cos BC

200
= 333 kN
0.6
200
BC = 100 + 266.7 =
= 366.7 kN
0.6
AC =

AC = 333 kN(C)

(7.3)
BC = 366.7 kN (T)

Due to the special loading condition at joint B , the force in member AB is zero.
2. Stress calculation: Using Eq. (7.1) we find the stresses in the members by dividing the member forces by their cross-sectional areas.
AC =

333 kN
50 cm

= 6.67

kN
= 66.7 MPa (C)(ans )
cm2

BC =

367 kN
30 cm

AB = 0(ans )

118

= 12.22

kN
= 122 MPa (T)(ans )
cm2

(7.4)

Example 7.2

A=3 in 2

3k

2
2 k A=1 in

3k

5k

A=2 in 2

2k

C 5k

Fig. 7.4 Example 7.2.


Problem: The bar in Fig. 7.4 is subjected to the forces shown. Find the axial stress in segments AB, BC and CD of the bar.
Solution:
1. Statics: We first find the magnitude of the internal axial force in the three segments of the bar. Cutting through the three
segments at an arbitrary location within the segment results in the three free-body diagrams shown below.

N AB

2k

FBD III

FBD II
3k

FBD I

N BC

2k

NCD

2k
D

3k B

Fig. 7.5 Free-body diagrams to find axial forces in segments of bar


Writing equilibrium in the horizontal direction for the three free-body diagrams gives
FBD I:

= 0 = 2 + N AB

FBD II:

= 0 = 2 6 + N BC N BC = 8 k

FBD III:

= 0 = 2 NCD

N AB = 2 k

NCD = 2 k

2. Stress calculation: The axial stresses in the three segments of the bar are thus
AB =

2k
1 in

= 2.00 ksi(ans )

BC =

8k
3 in

= 2.67 ksi(ans )

CD =

2 k
2 in2

= 1.00 ksi(ans )

We summarize the results by drawing the axial (normal) force and the axial stress diagrams for the bar.
8

8
2

2.00

2.67
2.00

+
A

+
C

[k]

2.67
+

C
1.00

1.00

119

[ksi]

Example 7.3

A=14 cm2

A=16 cm2

Cable 1
x

Cable 2

2.00 m

100 kN/m

10.00 m

Fig. 7.6 Example 7.3.


Problem: Two cables support beam as shown. Calculate the minimum and maximum distances x such that the axial stress in
each of the cables does not exceed the allowable stress of allow =10 kN/cm2 . Neglect the weight of the beam.
Solution: Since we know the cross-sectional areas of the cables and their allowable stress, we calculate the allowable forces
FA,allow = 14cm2 10

kN
= 140 kN
cm2

FB ,allow = 16cm2 10

kN
= 160 kN
cm2

(7.5)

We observe that the force in cable 1 reaches the allowable force when the distance x is at its maximum. Likewise, the force
in cable 2 reaches the allowable force when the distance x is at its minimum. The easiest way to obtain the two distances is
to use moment equations.

M
M

= 0 = 2 100 (10 x min 1) 140 10

x min = 2.00 m (ans )

= 0 = 2 100 (x max + 1) 160 10

x max = 7.00 m (ans )

x max
FA

2.00 m

x min

160 kN

100 kN/m

140 kN

10.00 m

2.00 m

100 kN/m

10.00 m

Fig. 7.7 Free-body diagram corresponding to x min and x max .

120

(7.6)

FB

7.2

Axial strain

7.2.1
Discussion
Forces acting on structural members cause either an elongation or a shortening of the members. Tensile forces or tensile
stresses lead to elongation of the members; compressive forces or compressive stresses cause the member to shorten. Since
the absolute elongation or shortening of the member provides little information unless we account for the length of the
member, engineers use the quantity of strain. Strain measures the elongation or shortening of a member relative to its original length. Strain caused by a shortening a member are called compressive strain, strain caused by elongation of the member are called tensile strain.
lengthening tensile strain
P

L + L
P

P
L

shortening compressive strain

L + L

Fig. 7.8 Illustration of strain.


The definition of axial or normal strain is thus

L
(Greek letter epsilon)
L

(7.7)

where L is the elongation (positive L ) or shortening (negative L ) of the member and L is the original (undeformed)
length. We observe that as the member elongates longitudinally it also contracts in the lateral direction. If a member shortens in the longitudinal direction, it expands laterally (see Fig. 7.8). This lateral displacement (or lateral strain) is commonly
much smaller than the longitudinal displacement (or longitudinal strain). In this class, we will only focus on longitudinal
strain. Note that strain is a dimensionless quantity which is typically very small (less than 0.5% for most building materials)
Example 7.4

cable 1

rigid

cable 2

3.00 m

2.00 m

7.2.2

= 0.5D
5.00 m

3.00 m

Fig. 7.9 Illustration of strain.

121

Problem: A rigid bar rotates about point A through an angle = 0.5D . Find the strain in cables 1 and 2 resulting from that
rotation.
Solution: We first calculate the elongation in the cables, i.e. the vertical displacements of points B and C . Since is a small
angle, we have
B = 0.5 /180D 5 m=0.0436 m

C = 0.5 /180D 8 m=0.0698 m

(7.8)

Using Eq. (7.7), we calculate the strains in the cables

1 =
7.3

0.0436 m
B
=
= 0.0145 = 1.45 %(ans )
3
3m

2 =

0.0698 m
C
=
= 0.0140 = 1.40 %(ans )
5
5m

(7.9)

Relation between stress and strain

7.3.1
Tension test
A tensile test, also known as tension test, is probably the most fundamental type of mechanical test we can perform on material. Tensile tests are simple, relatively inexpensive, and fully standardized. We conduct a tensile test by applying a gradually increasing tension force to a standardized cylindrical specimen (ASTM, American Society for Testing and Standards), which has a diameter of 0.5 inch and a length of 2 in (the length along which change in length is measured). We record both the elongation of the sample and the magnitude of the force. Based on the measured force and elongation we calculate stress and strain and obtain a stress-strain diagram. The stress-strain diagram provides important information about
the behavior and mechanical properties of the material.
We calculate the axial stress by dividing the load by the initial cross-sectional area of the sample.
=

P
A0

(7.10)

The use of the initial cross-sectional area to find the applied load is known as engineering stress or nominal stress. The
actual value of stress would require a continuous measure of the change in area of the samples cross-section. If the change
in cross-sectional area were measured and used to calculate the axial stress, then this would be known as the true stress or
the natural stress.
We find the strain in the specimen by dividing the measured elongation between the gauge marks by the initial gauge
length.
=

Testing
machine

L
L0

(7.11)

Extensometer

Specimen

Picture 7.1: Tension test setup.

122

7.3.2
Hookes law, yield strength, ultimate tensile strength
For tension tests of most materials, we will notice that in the initial portion of the test, the relationship between the applied
force and the elongation of the specimen is linear, that is stress is proportional to strain. The proportionality between stress
and strain discovered by Robert Hooke in 1676 is referred to as Hooke's Law
= E

(7.12)

where E (the slope of the stress-strain relation) is called the Modulus of Elasticity or Young's Modulus. Since the strain
is a dimensionless quantity, the modulus of elasticity has units of stress (ksi, MPa, etc.). For example, most types of steel
have a modulus of elasticity of E =29, 000 ksi or 210, 000 MPa . The modulus of elasticity of concrete varies between
E =3, 000 and 4,000 ksi, that of wood between E =1, 500 and 2000 ksi. The modulus of elasticity is one of the most important engineering parameters, widely listed in the literature. It describes the stiffness of the material, but it only applies to the
initial linear region of the stress-strain diagram. This region is called the elastic region. If we load a specimen within this linear region, the material will return to its exact same condition after removing the load. At the point where the curve deviates from the straight-line relationship, Hooke's Law no longer applies and some permanent deformation occurs in the
specimen. This point is called the elastic, or proportional, limit. Beyond this point, the material yields (no increase in
stress with increasing strain) and reacts plastically to any further increase in load or stress, i.e. it will not return to its original, unstressed condition after the load is removed. The upper yield stress is reached first, the stress than drops to a lower
yield stress. The upper yield strength or short yield strength y of a material is defined as the stress applied to the material at
which plastic deformation starts to occur while the material is loaded. The ultimate tensile strength u is the maximum stress
the specimen sustains during the test. It may or may not be identical to the strength f at failure. The difference between
maximum stress u and stress at failure f depends on the material. When the specimen reaches its ultimate stress u , the
cross-sectional area of the specimen reduces significantly, a phenomenon called necking.

u
f
y ,upper
y ,lower

elastic region ( = E )

Fig. 7.10 Qualitative stress-strain relation for steel.


7.4

Axial problems

7.4.1
Discussion
In this section, we will discuss a variety of problems involving axial stress, axial strain and axial displacements. Common to
all problems is that we use three fundamental relations learned before.
(1) Equilibrium (statics and definition of stress), (2) Strain-displacement relation, (3) Stress-strain-relation (HOOKEs law)
We first restrict our discussion to prismatic bars. A prismatic bar is a straight structural member whose cross section is constant along the length of the member. The shape of the cross section is arbitrary, e.g. circular, rectangular, I-shape, etc. As
already mentioned when discussing the concept of strain, axially loaded bars shorten under compressive forces and elongate
under tensile forces. We recall from Section 7.1, that we obtain the axial stress by dividing the axial force P by the crosssectional area A , hence

123

P
A

(7.13)

The linear stress-strain elation (HOOKEs law) is


= E

(7.14)

L
L

(7.15)

The strain-displacement relation is


=

Combining those three fundamental equations gives


L = L =

P
L =
L
E
EA

P = EA =

or

EA
L
L

(7.16)

for the elongation or shortening L , collectively referred to as deformation, of a prismatic bar. The deformation L is thus
proportional to the axial force P and the length L and inversely proportional to the area A of the cross section and the modulus of elasticity E of the member. We commonly refer to the product EA as the axial rigidity of the bar. In deriving the preceding equation, we have assumed that P, A and E are constant over the length L (see Fig. 7.11, Fig. 7.13a). Note that the
elongation or shortening of an axially loaded structural element (e.g. made of steel, aluminum, concrete, wood, etc.) is commonly small.
Modulus of elasticity E
Cross-sectional area A

E, A

P
L

L + L

Fig. 7.11 Deformation of a prismatic member.


It is instructive to recall from general physics the behavior of a simple spring in tension or compression (see Fig. 7.12. The
relation between the force F applied to a spring and the corresponding elongation L is given by
F = k L

(7.17)

where k is the stiffness of the spring, i.e. the value of force required to produce a unit elongation. From the preceding equation it becomes clear that the behavior of a bar under axial tension or compression is analogous to that of a simple spring,
since we can define the axial stiffness
k=

EA
L

(7.18)

of the bar in the same manner as for a spring.


L

L +L

Fig. 7.12 Analogy: Simple spring-axially loaded structural member


If P, A or E change abruptly, we can apply Eq. (7.16) to each segment of the bar in which the quantities are constant. The
displacement of the member then becomes (see Fig. 7.13b)
n

L =
i =1

Pi
L
Ei Ai i

124

(7.19)

If we allow P or A or both P and A to vary continuously over the length, we have to find L evaluating the integral
L

L =

P (x )

EA(x ) dx

(7.20)

An example for a situation in which both P and A vary along the length is a structural member with varying cross section
under its self-weight (see Fig. 7.13c).

(c)
(b)
(a)

Fig. 7.13 (a) Constant, (b) piecewise constant and (c) continuously varying cross section
7.4.2

Examples

Example 7.5
150 kN

100 kN

70 kN

B
2.00 m

4.00 m

20 kN

2.00 m

Fig. 7.14 Example 7.5.


Problem: Calculate the elongation of the steel bar AD having a cross-sectional area of A=5 cm2 due to the applied forces
shown ( E =200000 MPa ).
Solution: Using the concept presented in Example 7.2 we find the axial force N AB =100 kN, N BC = 50 kN, NCD =20 kN, for
the three segments. Using Eq. (7.19) then gives (using units of kN and cm)
LAD =

100
50
20
200
400 +
200 = 0.2 0.2 + 0.04 = 0.04 cm
5 20000
5 20000
5 20000

(7.21)

Example 7.6
10 in

250 k

5 in

E = 3000 ksi

x
7 ft

Fig. 7.15 Example 7.6.


Problem: Calculate the displacement of the end B of the circular bar AB . The diameter of the bar varies linearly from
d=10 in at A to d=5 in at B . Use E =3000 ksi .

125

Solution: We select units kips and inches and first express the diameter as a function of the coordinate x
d (x ) = 10

10 5
5

5 2
x = 10 x A(x ) = d 2 (x ) = 10 x
84
84
4
4
84

(7.22)

Since the cross-sectional area varies continuously, we use Eq. (7.20) and find
L

B =

P
dx =
EA(x )

84

250

5
3000 10 x
4
84

dx

(7.23)

84

1 1
= 1.783 = 0.178 in (ans )
5 10

4 250 84
1


3000 5 10 5 x
84

In solving the integral, we have used the identity


1

(a bx )

Example 7.7

dx =

1
b (a bx )

(7.24)

r0

r (x )

E,

E,

Fig. 7.16 Example 7.7.


Problem: Calculate the displacement of the end of the cone due to gravity. The material has specific weight .
Solution: The problem is similar to Example 7.6. The difference is that not only the area A but also the axial force P is a
function of x . This is because the axial force is the weight of the cone. We first express the radius as a function of the coordinate x
r
r (x ) = 0 x
(7.25)
L
Since the volume of the total cone is
V =

2
r0 L
3

(7.26)

the volume of a cone with a base radius r (x ) and length x is (see Fig. 7.16)

V (x ) = r (x )2 x
3

(7.27)

The cross-sectional area at height x is


A(x ) = r (x )2

Since P (x ) = W (x ) = V (x )

126

(7.28)

7.5

P (x )

dx =
EA(x )
E

V (x )

A(x ) dx = 3E
0

1 L2
(ans )
E

x dx = 6
0

(7.29)

Statically indeterminate axial force problems

7.5.1
Discussion
For all our problems in the preceding chapters we have been able to find the internal forces in a member by using the
equations of equilibrium. As we discussed in Chapter 3, structures of this type are termed statically determinate. If the number of unknown forces exceeds the number of independent equilibrium conditions, the structure is called statically indeterminate. An illustrative example of statical indeterminacy is a reinforced concrete column. Reinforced concrete is a composite material consisting of concrete and steel bars that reinforce the concrete. In conventional reinforced concrete design,
we assume that concrete and steel are rigidly bonded, i.e. both materials experience the same deformation. While we can use
statics to find the total force carried by the column, we cannot calculate the deformation of the column unless we know
the individual forces resisted by the concrete and steel bars, respectively. The additional equation needed for the solution is
a condition considering the geometry of the deformation. Commonly, such condition is referred to as a compatibility condition. Since we assume rigid bond between concrete and steel, the compatibility condition is that the deformations in steel
and concrete must be the same, hence
S = C =

(7.30)

Now we can express each deformation by the force-displacement relationship in Eq. (7.16).
S =

FS
L
AS ES

C =

FC
L
AC EC

(7.31)

such that
FS = AS ES

S
L

C
L

FC = AC EC

Using statics, we have

vert

= 0 =FS + FC P =

P
P

(AS ES + AC EC ) P
L

s = c

Ec , Ac
Es , As
Fs
2

Fs
2

Fc
Fig. 7.17 Reinforced concrete column.

127

(7.32)

7.5.2
Examples
Example 7.6

Area of steel:
Steel modulus:
Concrete modulus:

20 ft

AS = 4 in2
ES = 29, 000 ksi
EC = 3, 000 ksi

15 in

15 in

Fig. 7.18 Example 7.6.


Problem: If the shortening of the column in Fig. 7.18 is = 0.5 in , find (a) the force P acting on the reinforced concrete
column, (b) the compressive stress in both the concrete and the steel, (c) the percentage of the force P carried by the steel.
Solution: The force P is the sum of the forces carried by steel and concrete. The equation of equilibrium is thus

vert

= 0 =FS + FC P

(7.33)

The preceding equation, which is the only useful equation of equilibrium, contains two unknown forces. The structure is
thus statically indeterminate. Since we assume that concrete and steel shorten by the same amount
S = C = = 0.5 in

(7.34)

we can write

0.5
4 29000
=
= 241.7 k
20 12
L

0.5
=
FC = AC EC
(152 4) 3000 = 1381.3 k
20 12
L

FS = AS ES

(7.35)

The force applied to the column is


P = FS + FC = 241.7 + 1381.3 = 1623 k (ans )

(7.36)

The stresses in steel and concrete are


S =

FS
241.7
=
= 60.4 ksi(ans)
4
AS

C =

FC
1381.3
= 2
= 6.25 ksi(ans)
15 4
AC

(7.37)

The share of the steel in supporting the force P is


FS
241.7
=
= 14.9 %(ans)
1623
P

128

(7.38)

Example 7.7

ES , AS
EC , AC

1.00 m

Fig. 7.19 Example 7.7.


Problem: Twenty steel rods reinforce a concrete column as shown in Fig. 7.19. Find the diameter of each steel bar, if steel
and concrete carry 25% and 75% of the load, respectively EC = 30, 000 MPa, ES = 210, 000 MPa .
Solution: Since
FC =

AC EC = 0.75 P
L

FS =

AS ES = 0.25 P
L

(7.39)

and
A=

1.002 = 0.7854 m2
4

(7.40)

we have
FC
AE
(0.7854 AS )EC
= C C =
=3
FS
AS ES
AS ES

(7.41)

Solving for As yields


As =

0.7854Ec
0.7854 30000
=
= 0.0357 m2
3Es + Ec
3 210000 + 30000

(7.42)

The required diameter is thus


dreq =

4 0.0357
= 0.048 m = 4.8 cm (ans )
20

(7.43)

Note that the portion of the force carried by concrete and steel, respectively, is independent of both the column height and
the axial force applied to the column.

129

Example 7.8

2.00 m

3.00 m

rigid

B
5.00 m

3.00 m

Fig. 7.20 Example 7.8.


Problem: Two steel rods each having a diameter of 1 cm support a rigid beam that is pinned at A and subjected to a triangular distributed load with maximum intensity w . If the displacement of C is 1.5 cm downward, find the intensity w of the distributed load. Take Es =200, 000 MPa .
Solution: Given the displacement at C and knowing that beam is rigid, we can find the displacement at B by simple proportion (similar triangles) (see Fig. 7.21)
B = C
3.00 m

5
5
= 1.5 = 0.9375 cm
8
8

(7.44)

5.00 m

Fig. 7.21 Relation between displacements at B and C .


Next, we calculate the forces in the two steel rods
FB = B E A =
FC = C E A =

B
LB
C
LC

E A =

0.009375
200, 000 1 104 = 0.0625 MN
3

E A =

0.015
200, 000 1 104 = 0.0600 MN
5

(7.45)

Finally, we use the equilibrium condition to solve for the intensity w

=0

1
2
= FB 5.00 + Fc 8.00 w 8.00 8.00 = 0.0625 5.00 + 0.06 8.00 w 21.33
2
3
MN
kN
= 0.0371
= 37.1
m
m
FB

FC
w

A
FA

5 ft

3 ft

Fig. 7.22 Example 7.8. Free-body diagram.

130

(7.46)

Example 7.9

z
y

x
10

11

12

2m

2m

My

2m

2m

2m

Fig. 7.23 Example 7.9


Problem: Twelve vertical piles support a rigid pile cap against vertical displacements (see Fig. 7.23). The structure also contains diagonal piles for stabilization against lateral displacements. In this problem, we ignore the diagonal piles. All piles
have the same cross sectional area A= 400 cm2 , length L = 5 m and are made of concrete ( E =30, 000 MN/m2 ). If an
applied moment M y causes the pile cap to rotate by an angle of y = 1D about the y - axis through the center of the pile cap,
find the forces in the piles. Also find the moment M y . Clearly indicate whether the piles are in tension or compression.
Solution: Since only three vertical piles are necessary to stabilize the structure for vertical loads, the problem is statically
indeterminate. We thus need to express the force in each pile in terms of the corresponding pile deformation L . From the
given rotation of the pile cap, we can directly calculate the vertical displacement (deformation) of each pile (see Fig. 7.24)
1
2
3
4

= 5
= 6
= 7
= 8

= 9 = 3 m 1 /180
= 10 = 1 m 1 /180
= 11 = 1 m 1 /180
= 12 = 3 m 1 /180

= 0.0524 m
= 0.0175 m
= 0.0175 m
= 0.0524 m

1.00 m
3m

(piles 2,6,10)

(piles 1,5,9)

(lengthening)
(lengthening)
(shortening)
(shortening)

(7.47)

1.00 m
(piles 3,7,11)

3m
(piles 4,8,12)

2.00 m

1.00 m 1.00 m

2.00 m

Fig. 7.24 Relation between rotation of pile cap and displacement of each pile
From the lengthening or shortening of each pile, we calculate the pile forces from Eq. 7.16 (we use units of m and MN)

F1 = F5 = F9 =

EA
30, 000 400 104
1 =
0.0524 = 12.57 MN (tension)
L
5

F2 = F6 = F10 =

EA
30, 000 400 104
2 =
0.0175 = 4.19 MN (tension)
L
5

EA
30, 000 400 104
F3 = F7 = F11 =
3 =
0.0175 = 4.19 MN (compression)
L
5
F4 = F8 = F12 =

EA
30, 000 400 104
1 =
0.0524 = 12.57 MN (compression)
L
5

131

(7.48)

Note that since all piles have the same modulus of elasticity E , cross-sectional area A and length L , the pile forces in Eq.
(7.48) are proportional to the pile displacements in Eq. (7.47).
My
z

12.57

4.19

4.19
2.00 m

1.00 m 1.00 m

12.57

2.00 m

Fig. 7.25 Equilibrium between pile forces and applied moment


Expressing moment equilibrium for the free-body diagram in Fig. 7.25, we obtain the moment M y that is in equilibrium with
the pile forces

M = 0 = M

(12.57 3.00 + 4.19 1.00) 2 3 M y = 251MNm (ans )

(7.49)

Example 7.10

50 kN
C

2.00 m

2.00 m

4.00 m

Fig. 7.26 Example 7.10.


Problem: A rigid bar is suspended from three rods supporting both tension and compression. Find the forces in the rods for
the given loading. Each rod has the same modulus of elasticity E , cross-sectional area A and length L .
Solution: Since only two rods are necessary to stabilize the bar against vertical displacements, the problem is statically indeterminate. The free-body diagram for the structure and the two equations of equilibrium are shown below.

FA

50 kN

2.00 m

FB

2.00 m

FC

4.00 m

Fig. 7.27 Free-body diagram.

F = 0 = F + F + F 50
M = 0 = F 4 + F 8 50 2
y

(1)

(7.50)

(2)

The two equations of equilibrium contain three unknowns.

132

Since the beam is rigid, we can express the displacement of the structure in terms of two independent coordinates, say the
vertical displacement at B and the rotation . The displacement of any point along the bar is a function of those two coordinates.

4 A

C +

= A
A

4
4.00 m

4.00 m

4.00 m

4.00 m

B
4.00 m

4.00 m

Fig. 7.28 Vertical displacement as superposition of uniform displacement and rotation.


Measuring positive upwards, we obtain from Fig. 7.28
A = 4

B =

C = + 4

(7.51)

We can now express the unknown forces in terms of two independent displacement and have thus reduced the number of
unknown from three to two.

EA
EA
A =
( 4)
L
L
EA
EA
FB =
B =

L
L
EA
EA
FC =
C =
( + 4)
L
L
FA =

(7.52)

Note that according to our sign convention on the free-body diagram a positive displacement (upwards) causes a negative
force in the rods (compression). Likewise, a negative displacement (downwards) causes a positive force in the rods (tension). Substituting Eq. (7.52) into Eq. (7.50) yields

EA
EA
50
(1)
( 4 + + + 4) 50 = 3
L
L
EA
EA
M A = 0 = L (4 + 8 + 32) 100 = L (12 + 32) 100 (2)

=0=

(7.53)

Solving the previous two equations, gives


=

50 L
100 L
, =
3 EA
32 EA

(7.54)

Finally, we substitute the results for and into Eq. (7.52) and obtain

EA
100
50
( 4) = 4

3
L
32
EA
50
FB =
=
3
L
EA
100
50
FC =
( + 4) = + 4

3
L
32

FA =

= 29.167 kN(ans )
= 16.667 kN(ans )
= 4.167 kN(ans )

Fig. 7.29 summarizes the results by drawing a free-body diagram.

29.17 kN
2.00 m

50 kN
2.00 m

4.167 kN

16.67 kN
4.00 m

Fig. 7.29 Free-body diagram with known member forces.

133

(7.55)

Problems
7.1
z
z

2k
x

1k

5k
C

1 in 1 in 1 in 1 in

2k

Find the axial stress in sections AB and BC of the circular shaft.


Solution: AB = 0.318 ksi, BC = 0.398 ksi
7.2
2A

cable 1
x

cable 2
P
L

Determine the position x of a force P such that the tensile stresses in the two cables are equal. Cable 2 has a cross-sectional
area double that of cable 1.
Solution:

x
2
=
L
3

7.3

4 ft

2 ft

w
A

3 ft

Calculate the intensity w of the distributed load that can be supported if the allowable axial stress in the 0.5 in-diameter rod
BC is allow =22 ksi .
Solution: w = 0.576 k/ft

134

7.4

F
12 ft

15 k

D
12 ft

10 k

B
16 ft

For the truss shown, find the axial stress in member BC . The cross-sectional area of that member is 3.125 in 2 .
Solution: BC = 10 ksi(C)

cable 1

cable 2

B
5 ft

3 ft

2 ft

7.5

C
3 ft

A pin at A and two cables support a rigid bar:


(1) If loads acting on the bar cause point C to displace downwards c =0.5 in , find the strain in the two cables.
(2) If cable 1 has a stain of 1 =0.5 % , find the displacement of point C . Also find the strain in cable 2.
(3) If the bar rotates = 0.5D , find the strain in the two cables.
Solution: (1) 1 = 0.868 %, 2 = 0.833 %, (2) C = 0.288 in, 2 = 0.480 %, (3) 1 = 1.454 %, 2 = 1.396 %
7.6

2.00 m

2.00 m

B
3.00 m

C
(1) If point C of the truss displaces 1.30 cm downward, find the strain in truss members AC and BC .
(2) If point C of the truss displaces 1.30 cm downward and 2.00 cm to the right, find the strain in truss members AC and
BC .
Solution: (1) AC = BC = 0.0030 = 0.30 % (2)AC = 0.00608 = 0.608 %, BC = 0.0000550 = 0.0055 %

135

E, A

rigid

P,

7.7

A tension rod stabilizes a rigid post as shown. (a) For given values E , A, L, , derive a parameter k such that
P = k . (b) Find the angle that maximizes k .
Solution: (a) k =

EA
cos2 sin (b) 35D
L
6 ft

7.8

6 ft

9 ft

C
P

The truss is made of steel ( E =29, 000 ksi ) with a cross-sectional area of A=10 in2 . If point C of the truss experiences a
downward displacement C =0.20 in , find the magnitude of the member forces AC and BC and the magnitude of the
applied force P .
Solution: FAC = FBC = 372 k

P = 619 k

7.9
P

9 ft

B
12 ft

12 ft

The truss is made of steel ( E =29, 000 ksi ) with a cross-sectional area of A=10 in2 . What is the maximum allowable force
P if the horizontal displacement at B is limited to 0.58 in?
Solution: P = 876 k

136

20

4 6

7.10

6 4

20

[cm]

15 2

15

A column, whose cross section is shown above, is composed of a steel I-shape ( ES =210, 000 MPa ) encased in concrete
( EC =30, 000 MPa ). (a) If the column is subjected to a force P find the percentage of the force P carried by the steel. Neglect the circular fillet areas when calculating the area of the steel shape. (b) If the column experiences a strain of =0.3 % ,
find the force P applied to the column.
Solution: (a) Steel carries 53.3% of the load, i.e. FS = 0.533 P, FC = 0.467 P ; (b) P = 39.74 MN
y

7.11
4

10
8

The figure above shows in plan view a rigid pile cap supported by ten vertical piles. If a positive moment about the y - axis
is applied to the pile cap, identify the pile(s) (1) with the largest tension force, (2) with the largest compression force, (3)
with the smallest tension force, (4) with the smallest compression force, (5) with zero force. Repeat the problem for a moment applied about the x - axis
7.12
z
y

4
5

3
2

E , A, L
My

10
8

E , A, L
2.50 m

137

Ten vertical piles support a rigid circular pile cap against vertical displacements. The structure also contains diagonal piles
for stabilization against lateral displacements. In this problem, we ignore the diagonal piles. All piles have the same cross
sectional area A= 400 cm2 , length L = 5 m and are made of concrete ( E =30, 000 MN/m2 ). (a) If the pile cap is subject to a
moment M y =10 MNm , find the force in each pile. Clearly indicate whether the piles are in tension or compression. Also
find the rotation y of the pile cap due to the applied moment. (b) Repeat the problem for M x =10 MNm .
Solution:
(a) for M y
Pile
Force [kN]
(b) for M x
Pile
Force [kN]

1
800

1
0

2
647

3
247

2
470

3
761

4
247

5
647

4
761

6
800

5
470

6
0

7
647

8
247

7
470

8
761

9
247

9
761

10
647

10
470

Note: = compression,+ = tension, y = 0.0764D = 0.00133 rad

5 ft

7.13

A = 1 in2

A = 1in2

3 ft

A = 2 in2
1

5 k/ft

1 ft

2 ft

1 ft

2 ft

2 ft

4 ft

A rigid bar is suspended from three rods capable of supporting both tension and compression. Find the forces in the rods for
the given loading. Each rod has the same modulus of elasticity E .
Solution: F1 = 6.216 k, F2 = 2.568 k, F3 = 1.216 k (Tension)

E , A, L
E , A, L

E , A, L

rigid

F =100 kN

1.00 m 1.00 m

y
4

1.50 m

E , A, L

1.50 m

7.14

F =100 kN
2.00 m
2.50 m

2.00 m
2.50 m

A rigid slab is suspended from four struts capable of resisting both tension and compression. Each strut is made of the same
material and has the same cross-sectional area and length. A force F acts as shown. Find the forces in the four struts.
Solution: F1 = 93.75 kN, F2 = 31.25 kN, F3 = 43.75 kN, F4 = 18.75 kN
Note: = compression,+ = tension
138

7.15

t = 1 ft

2 k/ft

2 k/ft

C
3

10 ft

B
2
15 ft

20 ft

Three vertical concrete columns ( E = 3, 000 ksi ) support a 1-ft thick concrete slab as shown. In addition to the weight of
the slab the columns support a distributed load as shown. Find the areas of the columns such that each column shortens by
0.125-in. The specific weight of concrete is C =0.150 k/ft3 .
Solution: A1 = 12.0 in2 , A2 = 11.2 in2 , A3 = 13.6 in2
7.16

ES = 29, 000 ksi

10 in

A 10-ft high steel column has the tubular cross-section shown (outer radius = 10 in) and carries an axial compression force
of P = 2000 k . If the allowable stress of the material is allow =20 ksi and the allowable shortening of the column is
L =0.18 in , find the required thickness t of the tube.
Solution: t = 1.74 in

139

You might also like