You are on page 1of 21

Chapter 1

Internal Forces and Moments

When we construct a Free Body Diagram of an engineering system and


analyze the equilibrium state of that body, we are examining only the forces
and moments acting on the isolated body. However, for engineering design,
we eventually need to examine how loads are carried within a structure so
we can evaluate if it will fail or not. Within this chapter, we will examine this
process of evaluating internal loads and look at ways of visualizing internal
loading distributions. In fact, you have technically already done this for one
type of structure – a truss structure. We will build upon the internal force
analysis you have already performed for truss structures and look at internal
loading from a more general perspective.
10 kN
D E F
1.1 Concept of a Load Path
What is the purpose or function of a structure? In simple terms, you can L
define a structure as a construction designed to carry and transmit loads from A B C
one point to another. When we create a Free Body Diagram of a structure, we
define the origin of these loads as applied forces and moments (the action L L
forces of the surroundings on the structure) and the destination of these
loads as the support reaction forces and moments. The manner in which
10 kN
these loads are transmitted through the structure is known as the Load Path. FBD
The best way to understand the concept of a load path is to examine a
few simple structures. First, let’s consider two truss structures of the same
overall shape and size, but with different applied loads as illustrated in Fig.
1.1 and Fig. 1.2. Using the Free Body Diagrams provided in these figures, we Ax
can easily calculate the reaction forces for both load cases using equilibrium. y
For Load Case 1, we obtain: x
Ay Cy
X →+
Fx : Ax = 0 Figure 1.1 Truss load case 1 with
X ↑+ FBD.
F y : A y +C y − 10 kN = 0
X cc w+ ¡ ¢
M A : − (10 kN ) (L) + C y (2L) = 0

1
2 Chapter 1 Internal Forces and Moments

5 kN 5 kN
D E F Solving this system of equations for the unknown reaction forces, we obtain:

A x = 0 kN
L
A y = 5 kN
A B C
C y = 5 kN
L L
If we repeat the same equilibrium analysis for Load Case 2, we obtain:
X →+
5 kN 5 kN Fx : Ax = 0
FBD
X ↑+
F y : A y +C y − 5 kN − 5 kN = 0
X ccw+ ¡ ¢
M A : − (5 kN ) (2L) + C y (2L) = 0
Ax
Solving this system of equations for the unknown reaction forces, we obtain:
y
x
Ay Cy A x = 0 kN

Figure 1.2 Truss load case 2 with


A y = 5 kN
FBD. C y = 5 kN

Thus, the two load cases are statically equivalent as they produce the same
set of reaction forces on the body. One should be cautious, however, not to
confuse static equivalency with outright equivalency. The two load cases may
have the same reaction forces, but how the load is carried and transmitted
by the structure is very different. This becomes quite clear if we calculate the
internal forces in each of the two-force members of the truss structure for
each load case and overlay this information onto the truss.
Considering the first load case, we can easily identify that members AD,
DE , B E , E F , and F C all have to be zero-force members. Using the method
of joints at joint A and recognizing
p the problem
p is symmetric, we can easily
determine that N AE = −5 2 kN , NEC = −5 2 kN , and NE B = 10 kN , where
positive forces indicate tension and negative forces compression. For the
second load case, we can similarly identify that all the members except AD
and FC are zero-force members and that N AD = NFC = −5 kN . These results
are illustrated in the Fig. 1.3, with zero-force members shown by dashed
lines.

10 kN 5 kN 5 kN
FBD FBD

N −5
2k 2k
−5kN
−5kN

−5 N

y 10kN y
x x
5 kN 5 kN 5 kN 5 kN
Load Case 1 Load Case 2

Figure 1.3 Visualization of the load path within two statically equivalent
trusses.
1.2 Load Paths in Beams 3

Although both load cases are statically equivalent, the way the loads are
transmitted through the structure to the reacting supports is very different!
The two load cases produce different load paths through the structure.
The analysis of load paths can become more complex for other structures,
as we will see later on, but for now, this simple truss example illustrates
the concept well. We can even identify the usefulness of identifying and
analyzing such load paths for the design of such structures. We identified
earlier that the purpose of a structure is to carry and transmit loads, but as
structural designer, we need to size portions of the structure such that they
do not break or excessively deform under the loads they are transmitting. For
example, in the simple truss example above, the magnitudes of the internal
loads and the number of the truss members carrying the load are both higher
than in load case 2, suggesting that stronger and/or stiffer truss members
are required for load case 1.

1.2 Load Paths in Beams


w
Load paths within beams can quickly become complex due to the ability of
beams to carry multiple internal loads, including normal forces, transverse
A B
shear forces, bending moments, and torsional moments. Furthermore, these
loads can vary along the length of the beam. To illustrate this, consider a L
simple cantilever beam with a uniform distributed load, w, acting along the FBD
length of the beam as illustrated in Fig. 1.4. For the FBD given in the same w
MA
figure, we can solve for the reaction forces from the fixed support at A using
equilibrium: Az z

X →+ Ay L
Fz : Az = 0
y
X ↑+
F y : A y − wL = 0 Figure 1.4 Cantilever beam with a
X cc w+ L uniform distributed load.
M A : −M A − wL = 0
2
which leads to the result that:

Az = 0
A y = wL

wL 2
MA = −
2
This gives us the reaction forces imposed on the beam by the fixed sup-
port, but not the internal forces within the beam. These internal forces can
be revealed by sectioning the beam at various locations along its length, such
as every L/4 as illustrated in Fig. 1.5.
Please note that although in general a beam can have internal torsional
moments and normal forces, we can easily identify for this problem that the
given loading will not generate internal torsional moments or normal forces.
Thus, these particular internal forces and moments are omitted from Fig.
4 Chapter 1 Internal Forces and Moments

1.5. Additionally, the direction in which the internal forces and moments are
drawn in the FBD do matter. We will revisit this later in this chapter, but for
now, simply recognize the consistent manner in which the internal forces
and moments are drawn on each segment.

wL2 w M1 M1 V1 w M2 M2 V2 w M3 M3 V3 w

2
z

V1 V2 V3
wL L 4 L 4 L 4 L 4
y

Figure 1.5 Segmented FBD of the cantilever beam.

If we analyze the right-most segment of the beam, we can calculate the


internal bending moment M 3 and internal shear force V3 using equilibrium:

L
µ ¶
X ↑+
F y : V3 − w =0
4
L L
µ ¶µ ¶
X ccw+
M A : −M 3 − w =0
4 8

which leads to the result that V3 = 14 wL and M 3 = − 32 1


wL 2 . If we perform a
similar analysis on the other segments, we can easily show that V1 = 34 wL,
V2 = 12 wL, M 1 = − 12
9
wL 2 , and M 2 = − 18 wL 2 . The internal load varies along
the length of the beam as it transmits the distributed load to the reacting
fixed support. We can visualize this variation by plotting the internal loads
as a function of position along the beam, z, as illustrated in Fig. 1.6.

wL 3
4 wL
1
2 wL
1
1
4 wL
2
3 0
V(z) z
1
4 L 1
2 L 3
4 L L

1
4 L 1
2 L 3
4 L L
M(z) 3 z
2
2 0
1
− 321 wL
− 18 wL2
− 329 wL2
− 12 wL2
Figure 1.6 Internal load diagrams of the cantilever beam.
1.3 Sign Convention for Internal Forces and Moments 5

Although we have only calculated the internal loads at the indicated


discrete points for this problem, it is possible formulate precise equations
for the internal load distributions. For this particular problem, the internal
shear force distribution is described by V (z) = w (L − z) and the internal
moment distribution by M (z) = − w2 (L − z)2 . Do not worry about how to
determine these equations at the moment. How to do this will be covered
later in this chapter. For now, simply observe the potential complexity in the
internal load distributions and the implication they may have for design. For
instance, if we consider this cantilever beam as a simple model for an aircraft
wing with a distributed lift (albeit, upside down due to the distributed lift
acting downwards), we can observe the the highest internal loads are at the
fixed support end of the beam. The wing structure will thus have to have a
higher strength in this region.

1.3 Sign Convention for Internal Forces and Moments


Establishing a consistent sign convention is critical for effectively communi- P1 w1 P2
cating internal loads acting within a structure. For effectively communicating
the direction of forces and moments acting on a body, we have drawn free z
A B
body diagrams to establish know directions (in the case of applied loads) and y C

assumed directions (in the case of reactions) that can be used to decipher (a) Cantilever beam with
the meaning of the sign of the magnitude of forces and moments in our arbitrary loading
calculation. For instance, if the magnitude of a reaction force is calculated to Mx Mx
be negative, it means that it acts in the opposite to the direction indicated Nz Nz
in the free body diagram. Drawing a free body diagram for every possible C
point within a structure, however, is impractical. So a consistent convention
Vy Vy
is desirable.
(b) Positive internal forces
Strictly speaking, a sign convention can be arbitrary as long as it is an
acting at section C
agreed upon and consistently used. This is where many textbooks stop and
L=0
simply establish a convention and move on, but this is potentially dangerous Mx Mx
if a little though is not given to why internal loads would be communicated
Nz Nz
in an engineer sense. A structural engineer is responsible for ensuring a
Vy C Vy
structure is capable of transmitting the design internal loads without break-
ing (due to fracture, fatigue, or buckling) or excessive deformation. To do (c) Alternative view of positive internal
forces acting at section C
this, they will use the internal loads to calculate stresses (to evaluate when a
structure will break) and deformations. Thus, it is desirable to have a con- Figure 1.7 Sign convention for
vention that we can easily relate to these downstream uses of the internal positive internal force pairs.
loads.
Before we establish how the sign convention will relate to these uses,
let’s first establish what convention we will use. Fig. 1.7a shows a cantilever
beam with an arbitrary external loading that will produce internal normal
forces, shear forces, and bending moments. If we section this beam at an
arbitrary location, C , we can expose the internal forces and moments which
are drawn Fig. 1.7b in what we will use as our convention for positive. There
are two important things to recognize in the drawn internal forces. First,
our cut exposes action-reaction pairs of forces. We can emphasize this by
viewing the point where the internal forces are acting as a beam of zero
6 Chapter 1 Internal Forces and Moments

length as depicted in Fig. 1.7c. Second, the positive directions of these forces
and moments are related to the orientation of the coordinate system for the
problem. Don’t worry if the importance of these facts is not very obvious at
this point. Their importance will become more apparent when we discuss
the relationship between these internal forces and moments to deformation
and strength.

1.3.1 A Deformation-Based View on Positive Internal Forces


Mx Mx A critical application of internal forces in engineering is to analyses the
Nz
Nz deformation of a structure. It is thus logical to have a clear relationship
z
between how a structure deforms and the internal forces that cause that
y Vy Vy deformation. The relationship correlating to our sign convention for internal
(a) Positive internal force pairs forces to deformation is illustrated in Fig. 1.8.We will break down these
relationships one by one.
Nz First, we will start with internal normal forces, as they are the easiest
z Nz
and most intuitive to understand. We define a positive internal normal force
y as one that causes the structure to elongate along the line of action of the
(b) Elongation caused by force, as shown in Fig. 1.8b. Conversely, a negative internal normal force
positive normal force would cause a structure to become shorter (or contract) along the line of
action of the force. Normal forces are so ubiquitous in are daily lives that
dy we haven even come up with different vocabulary to differentiate between
= positive
dz
z
positive and negative normal forces, where we typically refer to positive
y
Vy internal normal forces as tension and negative internal normal forces as
Vy compression.
(c) Positive slope caused by Next, we will examine internal shear forces. If we examine the mode
positive shear force of deformation in shear illustrated in Fig. 1.8c, we see that shear causes a
change shape (causing the rectangular beam to skew) rather than an elon-
Mx Mx gation or contraction. As a result, we have to come up with another way to
concave

z
relate the sign of the force to the deformation. This can be done by looking
convex at the angle at which the beam skews as a result of the shear.
y ³ ´ We define a
dy
(d) Convex (positive) curvature on positive internal shear force as causing a positive slope d z with respect
positive side of beam caused by to the coordinate system being used for the internal forces. As y is positive
positive bending moment
downwards in Fig. 1.8, then the downwards sloping deformation of the beam
Figure 1.8 Positive deformations in the z-direction is considered positive.
resulting from positive internal Finally, we have the internal bending moment. Deformation due to
forces bending is more complicated than due to normal or shear forces as bend-
ing moments cause curvature. This curvature can cause both positive and
negative slopes in the beam, and can cause both elongation and contraction
on the convex and concave sides of the beam respectively. Mathematically
speaking,
³ we ´ can relate curvature to the second derivative of the deformed
d2y
shape d z 2 ; however, this tends to be less intuitive to the average engineer.
So instead, we can relate the sign of the moment to the location where elon-
gation occurs (ie: the location of the convex side of the beam). We will thus
define an internal moment as positive if it causes elongation of the side of
the beam facing the positive coordinate direction.
There is one internal load we have not considered in Fig. 1.8, which is
1.4 Internal Force Diagrams 7

an internal torque (a moment aligned with the axis of the beam). Similar
to internal shear force, an internal torque does not result in elongation or
contraction of the beam. It will result in twisting of the beam around its
central axis. If this angle of twist (which we will denote as φ) increases
(according to the right hand rule) in the positive coordinate direction along

the beam, then the torque is defined as positive. Conversely, if d z is negative,
the internal torque is also negative. This will become more apparent in later
Mx Mx
chapters on the topic of torsion. For now, we will focus on internal normal, Nz
shear, and bending. z Nz
y Vy Vy
1.3.2 A Stress-Based View on Positive Internal Forces (a) Positive internal force pairs

Internal forces do not actually exist as point forces within a structure. In


reality, these forces are distributed over an area where we refer to the intensity
x z Mx
of this distributed force at any given point in that area as a stress. These
y
internal forces are thus the resultant (or sum) of all of these stresses acting Nz
over the area they are acting on. It is thus logical that the sign convention of
internal forces relates to the sign convention of the internal stresses they are Vy
a resultant of. (b) 3D view of positive cutting plane
At this point of your studies, you may not have looked at in detail what
stress states are generated by each type of internal force or examined what
x z
the sign convention is for internal stresses. However, we can illustrate the
y
stress state generated on the positive cutting plane (plane with an outward +σzz
normal in the positive z-direction) as done so in Fig. 1.9. Here you can see
that a positive normal force will create a tensile (positive) normal stress on
(c) Positive Nz causes positive
this plane. Similarly, the positive internal shear force will result in a shear
(tensile) σzz
stress acting downwards in the positive y-direction on this cutting plane.
Finally, as an internal bending moment creates both tensile (positive) and
compressive (negative) normal stresses, we define the sign of the moment
x z
based on the sign of the normal stress in the positive coordinate of the
y
section. As shown in Fig. 1.9e), the sign of the normal stress matches the
sign of the y-coordinate within the section for a positive bending moment. +τzy
Don’t worry too much if this stress viewpoint of positive internal forces (d) Positive Vy causes positive τzy
is confusing at this time. We will revisit it and discuss its impact on the
equations we derive for internal stresses in later chapters. −σzz
x z
1.4 Internal Force Diagrams y
+y

In the previous section, we established a sign convention that we can use


+σzz
to effectively communicate internal loads within a structure for a given
(e) Positive Mx causes positive
coordinate system. Using this convention, it is common to communicate (tensile) σzz in the positive y
internal loads as plots of the internal load variation within the structure – half of the cross-section
the so-called Internal Force Diagrams. We have already encountered such
diagrams when we examined the load path within a simple cantilever beam Figure 1.9 Stresses resulting from
with a uniform distributed load by sectioning the structure at discrete points positive internal forces
(Fig. 1.4). This approach will work in general, but does not allow a precise
determination of the functions describing the internal force distributions.
8 Chapter 1 Internal Forces and Moments

To solve for the internal force distributions, we would like to set up the
equilibrium equations for the internal forces as functions of distance along
the beam (the coordinate distance z in the previous examples). In order to
do this, we will draw a parametric FBD exposing the internal forces within
the structure at an arbitrary distance along the beam. With this FBD, we can
use equilibrium to solve for the internal forces as a function of position along
the beam.
Let’s see how this would work for the uniform loaded cantilever beam we
examined in the previous section. The original problem with the parametric
FBD drawn below it is shown in Fig. 1.10. For now, we will leave the reaction
forces at point A as variables, but recall that we have already solved for
these in section 1.2. Recognizing the beam is in static equilibrium, we can
formulate equilibrium equations and solve for the unknown internal forces.
Applying force equilibrium in the z direction:
w
X →+
F z = 0 = −A z + N z
A B
⇒ Nz = A z
L
FBD Similarly, force equilibrium in the y-directions results in:
MA w
Mx X ↑+
Nz
Az F y = 0 = A y − w z − Vy
z
Vy ⇒ Vy = A y − w z
Ay z
y The internal bending moment can then be obtained from moment equilib-
rium. To eliminate the other internal loads from this expression, we will
Figure 1.10 Parametric FBD of the evaluate moment equilibrium at the location of the cut, z:
uniform loaded cantilever beam
X ccw+ ³z ´
M = 0 = −M A − A y z + w z + Mx
2
w z2
⇒ Mx = M A + A y z −
2
These results are in terms of the reaction forces at A as variables. If we take
the results for the reaction forces from section 1.2

Az = 0
A y = wL

wL 2
MA = −
2
We can obtain our final expressions for the internal forces as functions of
position z along the beam:

Nz = 0 (1.1)
V y = w (L − z) (1.2)
w
Mx = − (L − z)2 (1.3)
2
1.4 Internal Force Diagrams 9

These results have been plotted in Fig. 1.11.


w
Take a moment to look at the final results in equations 1.1 through 1.3
and Fig. 1.11. What is critical to interpreting what these results mean if you z

did not perform the calculations yourself? A B

The answer is the coordinate system. The sign convention for the internal y L
forces, and thus the meaning of the signs in our internal force diagrams, is 0
Nz
dependent on the orientation of the coordinate system. Furthermore, the
wL
mathematical description of the position along the beam, z, in equations 1.1
through 1.3 only makes sense if the origin and orientation of the coordinate +
Vy
system is known. We often use the same coordinate system position and
orientation, so it is easy to get into the bad habit of not clearly establishing
your coordinate system in your problem. Please do not get into this bad Mx
-
habit! It is a dangerous habit that can lead to miscommunication in your
engineering career. − 12 wL2
Another critical step in our analysis that is often overlooked is the fact that
the internal forces in our FBD in Fig. 1.10 were drawn in the positive sense Figure 1.11 Plotted internal force
with respect to our established coordinate system. The signs in equations 1.1 diagrams
through 1.3 are relative to the direction we drew the forces in our FBD. Thus,
it is good practice to draw them in the positive sense to ensure the solutions
we obtain for the internal forces using equilibrium and our FBD will have
the correct sign.
t Please make sure to always
clearly define the coordi-
1.4.1 Beams with Discontinuous Internal Force Distributions
nate system you use to con-
The example of a cantilever beam with a uniform distributed load exam- struct your internal force
diagrams. This is typically
ined in the previous section ended up having internal force distributions
done with a clear FBD.
that could be described by continuous mathematical functions. This was
convenient as it meant that a single parametric FBD could be used to create
a set of equilibrium equations to solve for the internal force distributions.
However, this will not always the case for all structures and load cases.
L/2 P = wL
To illustrate this, let’s examine the internal force distribution of a can-
tilever beam with a point force that is statically equivalent to the distributed
load in the previous section. For the point load to be statically equivalent, it A B
should have a magnitude of wL and act through the centroid of the uniform
L
distributed load, which would be mid-span of the beam, as illustrated in Fig.
FBD
1.12. As it is statically equivalent, it will have the same reaction forces as the P = wL
previous problem: MA L/2

Az z
Az = 0
A y = wL Ay L
y
wL 2
MA = −
2 Figure 1.12 Cantilever beam with
a mid-span point load
If we try and follow the steps from the previous problem, we would
now draw a parametric FBD that is valid for any value of z along the beam.
However, this is not possible in a single parametric FBD due to the presence
of the point load at z = L2 . Instead, we will have to create two parametric
FBDs to account for the influence of the point load as shown below. Please
10 Chapter 1 Internal Forces and Moments

note that the two ranges do not actually include the point z = L2 where the
point force acts. This is because we cannot actually make a cut through the
middle of a point force because a point has no width.
FBD: ( 0 < z < L 2 ) FBD: ( L 2 < z < L )
P = wL
MA Mx MA Mx
Nz Nz
Az z Az z
Vy Vy
Ay z Ay z
y y

Figure 1.13 Parametric FBDs for the two ranges 0 < z < L2 and L2 < z < L .
¡ ¢ ¡ ¢

t Do not be deceived into We can now examine equilibrium for each parametric FBD to determine
thinking that the influ- the internal forces acting within its relevant range of positions along the
beam. First, examining the range 0 < z < L2 :
¡ ¢
ence of the point load is
not present in both para-
metric FBDs in Fig. 1.13. X →+
F z = 0 = −A z + N z
Indeed, the values of the
(1.4)
L
µ ¶
reaction forces at point A
are dependent on the point ⇒ Nz = A z = 0 0<z <
2
force itself; thus, its effect is
present within the reaction X ↑+
forces themselves. F y = 0 = A y − Vy
(1.5)
L
µ ¶
⇒ V y = A y = wL 0<z <
2
X cc w+
M = 0 = −M A − A y z + M x
µ
L
¶ µ
L
¶ (1.6)
⇒ M x = M A + A y z = wL z − 0<z <
2 2
For the second range, we could apply the equilibrium equations to the
FBD (L 2 < z < L)
second parametric FBD in Fig. 1.13. However, it is actually much easier in
Mx this case to consider the portion of beam that is cut away in that parametric
Vy
A Nz
FBD. This would be a segment of beam of length (L − z) with no applied loads
z
or reaction forces acting on it as illustrated in Fig. 1.14. Using this alternative
z parametric FBD, it is trivial to see that:
y µ
L

Nz = 0 <z <L (1.7)
2
¡ Alternative
Figure 1.14 parametric
FBD for L2 < z < L .
¢
L
µ ¶
Vy = 0 <z <L (1.8)
2
L
µ ¶
Mx = 0 <z <L (1.9)
2

Plotting these results in Fig. 1.15, we can clearly see the discontinuity in the
internal force diagrams created by the point force. Such discontinuities will
be created by any change in the loading as you travel along the length of the
beam. Try and take note of how the loading on the beam introduces such
discontinuities in the following examples.
1.4 Internal Force Diagrams 11

L/2 P = wL

A B

y L
0
Nz

wL
+ 0
Vy

0
Mx
-
− 12 wL2

Figure 1.15 Internal force diagrams for cantilever beam with a mid-span
point load.

Example 1.1
Calculate and draw the internal normal force, shear force, and bending mo-
ment diagrams for a simply supported beam subjected to an offset point
force P acting at an angle θ illustrated in Fig. 1.16.

P
θ
A C
B

a b
L
Figure 1.16 Simply supported beam for Example 1.1

Solution: The question is asking us to calculate and plot internal force di-
agrams for the beam. To do this effectively, we need to have a coordinate
frame to base the signs of our internal forces on and we need to know the
magnitude and direction of the reaction forces at the two simply supported
ends of the beam. As these are not provided to us in the problem, we will
start at this point.
P
First, we will draw an FBD of the beam to define the assumed directions
A θ
of our reaction forces and establish a coordinate frame as illustrated in Fig. C
Az z
1.17. Here, we are using a coordinate frame that is centred on the left end B
of the beam, has a positive y-coordinate below the beam, and a positive Cy
Ay a b
z-coordinate along the length of the beam.
y L

Figure 1.17 FBD of the simply


supported beam
12 Chapter 1 Internal Forces and Moments

As the beam is in a state of static equilibrium, we can formulate equilibrium


equations to solve for the unknown reaction forces:
X →+
F z = 0 = A z − P cos θ (1.10)
X ↑+
F y = 0 = A y − P sin θ +C y (1.11)
X cc w+
M A = 0 = − (P sin θ) (a) +C y L (1.12)

Using equations 1.10 and 1.12, we can directly evaluate that

A z = P cos θ
Pa
Cy = sin θ
L
Substituting the result for C y into equation 1.11, we obtain
³ a´
A y = P sin θ 1 −
L

which can be further simplified by recognizing that 1 − La is equal to b


¡ ¢
L

Pb
Ay = sin θ
L
Mx
A Nz With the reaction forces now known, we can begin to solve for the internal
Az z force distributions. We do this by drawing a new FBD where we cut into the
beam at an arbitrary distance z along the beam as illustrated in Fig. 1.18. If
Ay z Vy
you take a moment to think about the applicability of this FBD to the problem,
you can easily identify that it is not valid for all values of z along the beam.
y
It is only valid for the range 0 < z < a. Limiting ourselves to this range of z,
Figure 1.18 Parametric FBD expos- we can formulate equilibrium equations to solve for the internal forces as
ing internal forces from 0 < z < a functions of z. Looking at force acting in the z-direction
X →+
Fz = 0 = A z + Nz
⇒ N z = −A z = −P cos θ (0 < z < a)

Similarly, for forces acting in the y-direction


X ↑+
F y = 0 = A y − Vy
Pb
⇒ Vy = A y = sin θ (0 < z < a)
L
Finally, for moments acting about point A
X cc w+
M A = 0 = Mx − Vy z
P bz
⇒ Mx = Vy · z = sin θ (0 < z < a)
L
P
Mx We can now repeat this analysis by drawing a FBD that is valid for a < z < L as
A θ Nz shown in Fig. 1.19 and repeating the equilibrium analysis as before. Looking
Az z at force acting in the z-direction

Ay z Vy X →+
F z = 0 = A z − P cos θ + N z
(1.13)
y
⇒ N z = P cos θ − A z = 0 (a < z < L)
Figure 1.19 Parametric FBD expos-
ing internal forces from a < z < L
1.5 Internal Force (NVM) Relationships 13

Similarly, for forces acting in the y-direction


X ↑+
F y = 0 = A y − P sin θ − V y

⇒ V y = A y − P sin θ
b
µ ¶
= P sin θ −1 (1.14) (noting L − b = a ⇒ b
L − 1 = − La )
L
Pa
=− sin θ (a < z < L)
L
Finally, for moments acting about point A
X cc w+
M A = 0 = −P sin θ (a) − V y z + M x

⇒ M x = V y z + P a sin θ
Pa
=− sin θ · z + P a sin θ (1.15)
L
Pa
= sin θ (L − z) (a < z < L)
L
Summarizing the results, we are left with the following equations for the
internal force distributions as a function of z: P
A θ
C
Az z
N z = −P cos θ (0 < z < a) B
(1.16) Cy
Ay a b
=0 (a < z < L)
y L
Pb
Vy = sin θ (0 < z < a)
L
(1.17) Nz
Pa - 0
=− sin θ (a < z < L) − P cos θ
L
Pb
P bz sin θ
sin θ L Pa
Mx = (0 < z < a) − sin θ
L + L
Pa
(1.18) Vy
= sin θ (L − z) (a < z < L) -
L
Pab
sin θ
With the distributions now know, we can answer the second part of the L
question by drawing the internal force diagrams. Fig. 1.20 shows the internal +
force distributions graphically based on equations 1.13 through 1.18. It is Mx
important to note that the signs for the internal forces in equations 1.13
through 1.18 are correct as we drew the internal forces in our FBDs in the Figure 1.20 Plots of internal force
positive sense. diagrams

1.5 Internal Force (NVM) Relationships


If you analyze and solve for the internal force distributions of enough beam
problems, you will may quickly come to realize two facts:

1. Solving for the internal force distributions using equilibrium can be-
come quite tedious, particularly when there are several discontinuities
in the distributions.
14 Chapter 1 Internal Forces and Moments

2. There seem to be certain patterns between the NVM diagrams that


seem to emerge, suggesting that there are some underlying relation-
ships between these diagrams and the loading that causes them.

The first point is not stated simply to be humorous or imply that students
dislike doing work. It actually illustrates one of the greatest motivators in
engineering —the feeling that there must be a better way to do something!
While, the second point highlights the beginnings of a solution to that feeling.
The identification of a pattern that hints at an alternative approach that
could be generalized to any loading.
In the remainder of this section, we will look at deriving the interrelation-
ships between internal and external loads acting on a beam. To do so, we
have to first generalize the loading that can occur on a beam. For the time
being, we will restrict ourselves to a 2D problem, resulting in the generalized
loading of a beam illustrated in Fig. 1.21. Please take note of the coordinate
system given in Fig. 1.21, as well as its orientation. The relationships we will
derive will be dependent on this coordinate system and the orientation of
the external loads with respect to it. For this reason, all external loads will be
considered positive when acting in the positive coordinate directions, which
is how they have been drawn below.
z

y
P1 P2
w(z)

A B
M1 M2

z z P2y
P1y
y y
wz(z) wy(z)

A
P1z P2z
B
+ A
M1 M2
B

Lateral Loads Transverse Loads

Figure 1.21 Generalized 2D external loading of a beam split up into lateral


and transverse external loads.

Figure 1.22 Infinitesimal element In addition to the generalized external loading, we also need to general-
of the beam located at distance z ize the internal loading as well. To do this, we need to consider an arbitrary
showing the variation of internal infinitesimal element of the beam of width d z located a distance z from the
loading across the element. left hand side of the beam, as illustrated in Fig. 1.22. All internal loads have
been drawn as positive for the coming derivations. With the internal and
external loads generalized, we can now evaluate the relationships between
them. This will first be done relatively quickly for the sake of brevity; how-
ever, if you require a more detailed explanation of the derivation, you are
encouraged to review the video derivation available through the QR-code in
Fig. 1.23.
1.5 Internal Force (NVM) Relationships 15

1.5.1 Relationships for Lateral External Loads


Take a moment to consider the infinitesimal element shown in Fig. 1.22 and
the generalized lateral loading in Fig. 1.21. How do you expect externally
applied lateral loads on a beam to affect the internal loading? Do you expect
the lateral loads to cause bending? How about shear? And what about the in-
ternal normal forces? Try visualizing this by thinking about the deformation
you would expect in the beam as a result of the lateral loading.
Remember, a beam is a one-dimensional model of a structure, so the
externally applied lateral loads are in line with the centre of the beam. Lateral
loads will thus only cause elongation (tension) or contraction (compression)
of the beam along its length. We would not expect bending or shear to occur.
Knowing this, we can simplify the infinitesimal element we examine for
lateral loads by only considering the internal normal forces. If you are unsure
about this simplification, it is proven in the video derivation accessible
through the link in Fig. 1.23.

Lateral Point Force

Let’s first consider the effect of a lateral point force on the internal loading
within a beam. Fig. 1.24 shows an infinitesimal element cut from a beam at
the location where an externally applied point lateral force is acting. We can
examine the influence of this force on the internal normal force by evaluating
equilibrium of the element in the direction of the normal forces.
X →+
F z = 0 = −N z + P z + (N z + d N z )
⇒ d N z = −P z

Here we see that there is no dependence on the width of the element, d z, Figure 1.24 Infinitesimal element
thus, in the limit as d z approaches zero, we obtain: with a lateral point force.

lim d N z = ∆N z = −P z (1.19)
d z→0

This results indicates that a step change in the internal normal force, ∆z,
occurs at the location where a point lateral force, P z , occurs. The negative
sign simply indicates that for our sign convention for positive internal and
positive applied forces, a positive applied force would cause the internal
normal force to become more compressive (negative) when moving along
the beam in the positive z direction.

Lateral Distributed Force

Next, we will consider the effect of a lateral distributed force, w z (z), acting
along a beam. The generalized distributed force is a function of position z
along the length of the beam, but when examining an infinitesimal element
of the beam, we have to remember that we are effectively evaluating a point
along the beam. Thus, the distributed force has a constant value equal to

Figure 1.25 Infinitesimal element


with a lateral distributed force.
16 Chapter 1 Internal Forces and Moments

the distributed force at that point as illustrated in Fig. 1.25. Evaluating


equilibrium of forces along the length of the beam, we obtain:
X →+
F z = 0 = −N z + w z d z + (N z + d N z )
(1.20)
d Nz
⇒ = −w z
dz
Here we see that we obtain a differential equation for the internal normal
force. The rate of change of the internal normal force, ddNzz , is simply the
slope of the internal normal force diagram.

1.5.2 Relationships for Transverse External Loads


As with the lateral loads, we can first stop to think about how we expect
transverse loads to influence the internal loads within a beam. Looking at
the transverse loads illustrated in Fig. 1.21, how would you expect these
types of loads to influence internal bending, shear, and normal forces within
the beam?
For the lateral loads, we saw that the variation in internal normal force
was cause by a disturbance introduced by the externally applied loads that
acted parallel to the beam. As none of the transverse loads act parallel to
the beam, we do not expect they will disturb the equilibrium of the internal
normal forces. We can thus limit our examination of transverse loads to their
effect on internal bending moments and shear forces. If you are unsure about
this simplification, it is proven in the video derivation accessible through the
link in Fig. 1.23.

Transverse Point Force

The infinitesimal beam element with a transverse point force acting on it is


given in Fig. 1.26. Evaluating force equilibrium in the vertical direction, we
obtain:
X ↑+ ¡ ¢
F y = 0 = V y − P y − V y + dV y
⇒ dV y = −P y

Taking the limit of this result as d z approaches zero, we obtain:

lim dV y = ∆V y = −P y (1.21)
d z→0

Figure 1.26 Infinitesimal element For moment equilibrium, we will select a point A located in the cen-
with a transverse point force. tre of the element such that the point force P y does not contribute to the
expression:
X cc w+ dz ¡ ¢ dz
M A = 0 = −M x − V y − V y + dV y + (M x + d M x )
2 2
dz
⇒ d M x = V y d z + dVz
2
1.5 Internal Force (NVM) Relationships 17

Taking the limit of this expression as d z approaches zero, we obtain:

lim d M x = ∆M x = 0
d z→0

which means that there is no step change in the internal moment of a beam
due to the discontinuity of a transverse point force P y .

Transverse Distributed Force

The infinitesimal beam element with a transverse distributed force acting


on it is given in Fig. 1.27. As for the case of the lateral distributed force, this
distributed force can be considered as a constant value on the infinitesimal
element as the element represents a single point along the beam. Evaluating
force equilibrium in the vertical direction, we obtain:

X ↑+ ¡ ¢
F y = 0 = V y − w y d z − V y + dV y )
(1.22)
dV y
⇒ = −w y
dz
resulting in a similar relationship between the variation of internal shear
force and the transverse distributed force as found between internal normal
force and lateral distributed force.
Evaluating moment equilibrium around point A indicated in Fig. 1.27,
we obtain:
dz
µ ¶
X cc w+ ¡ ¢
M A = 0 = −M x − w y d z − V y + dV y d z + (M x + d M x ) Figure 1.27 Infinitesimal element
2
with a transverse distributed force.
(d z)2
0 = −w y − V y d z − dV y d z + d M x
2
To further simplify this expression, we can recognize that all of the infinites-
imal terms (ie: dV y , d M x , and d z) are very small numbers, so parts of the
expression with higher order terms of these quantities will be negligible. This
means that the term with (d z)2 and with d zdV will be negligible compared
to the terms with only d z and d M x . This leaves us with the expression:

0 = −V y d z + d M x
d Mx (1.23)
⇒ = Vy
dz

Point Moment Couple

The last load to consider is a point applied moment couple as illustrated in


Fig. 1.28. Here, we can recognize that the presence of the moment will not
alter force equilibrium, so we only need to evaluate moment equilibrium for
the element:
X cc w+ ¡ ¢
M A = 0 = −M x + M 0 − V y + dV y d z + (M x + d M x )
¡ ¢
⇒ d M x = −M 0 + V y + dV y d z

Figure 1.28 Infinitesimal element


with a point moment couple.
18 Chapter 1 Internal Forces and Moments

Taking the limit of this expression as d z approaches zero:

lim d M x = ∆M x = −M 0 (1.24)
d z→0

Thus, just as for the point forces, a point moment couple will cause a step
change in the internal moment distribution.

1.5.3 Applying Relationships to Construct NVM Diagrams


In order to understand how we can use the results of the previous section to
construct NVM diagrams, let’s first take a look at how they relate to an NVM
diagram we have already constructed in Example 1.1. For a reminder, the fi-
nal result for the NVM diagrams we obtained in that example are reproduced
in Fig. 1.29. We will look at each diagram one by one and demonstrate how
the results of eqns. 1.19 through 1.24 can be used to construct them.

Normal Force Diagram


P From all of the relationships that were derived between external loading and
A θ internal loading, only two of them relate to the internal normal force:
C
Az z
B d Nz
Cy = −w z
Ay a b dz
L
y ∆N z = −P z

Nz Recognizing that the N -diagram itself is a function, these two equations tell
- 0
us that:
− P cos θ
Pb • The slope of the N -diagram is equal to the magnitude of the intensity
sin θ
L Pa
− sin θ of the lateral distributed force
+ L
Vy
- • Discontinuous jumps in the N -diagram (∆N ) occur at lateral point
Pab forces equal to the magnitude of the point force
sin θ
L
One must be conscious of the signs in these equations. Lateral forces acting
+
Mx in the positive z-direction will result in internal normal forces becoming
more compressive when moving in the positive z-direction, hence the nega-
Figure 1.29 Final result from Ex- tive sign in both equations.
ample 1.1. Looking at the FBD and N -diagram from Example 1.1 below, we can
clearly see how these relationships could be used. Traveling along the FBD in
the positive z-direction, we can transform the loads on the FBD into the N -
diagram. We first encounter the lateral point force A z acting in the positive
z-direction at point A. At this location, a step change in the N -diagram
occurs in the negative direction as A y is itself positive. Continuing along
the beam, the internal normal force distribution has a slope of zero, as the
intensity of the lateral distributed force acting on the beam is zero. This
continues until another lateral point force equal to −P cosθ is encountered
at point B . As this force is acting in the negative z-direction, it causes a
positive jump in the internal N -diagram. Continuing along the beam, the
slope of the N -diagram remains zero as the intensity of the lateral distributed
force acting on the beam is zero.
1.5 Internal Force (NVM) Relationships 19

P
A θ
C
Az z
B
Cy
Ay a b

y
L

slope = 0
Nz
Az - P cos θ

slope = 0

Figure 1.30 Construction of an N -diagram graphically from the FBD.


This particular example is relatively simple as there is no lateral dis-
tributed force acting on the beam. However, we can rearrange our view of
the internal normal force relationships by looking at the normal force at any
point along the beam as a summation of the variations in internal normal
force up to that point along the beam:

Zz n
X
N z (z) = −w z d z + −P zi
i =1
0

where the integration and summation are both performed from the end
of the beam (z = 0) up to the point of interest, travelling in the positive
z-direction.

Shear Force Diagram

Looking at the relationships that were derived related to the internal shear
force, again we are left with only two that apply:

dV y
= −w y
dz
∆V y = −P y

This is analogous to the case for the normal force diagram, and we can
take the same graphical approach, only now we need to consider the trans-
verse loads rather than the lateral loads. The signs can become more trou-
blesome for transverse loads as the difference between positive and negative
shear is less intuitive then the difference between positive (tensile) and nega-
tive (compressive) normal force. To help deal with this, you may have noticed
that the y-direction in the FBDs of this chapter have been consistently drawn
with y as positive downwards. This benefits us now as it means the exter-
nally applied transverse loads point in the direction that they will change the
V -diagram. This can be clearly seen in the V -diagram below.
20 Chapter 1 Internal Forces and Moments

P
A θ
C
Az z
B
Cy
Ay a b
y L
slope = 0
Ay + P sin θ
Vy
- Cy

slope = 0

Figure 1.31 Construction of a V -diagram graphically from the FBD.


This example, like the N -diagram, is relatively straight forward as there
are no transverse distributed loads. We can still, however, rearrange our
view on the internal shear force relationships by looking at the beam as a
summation of infinitesimal elements:
Zz n
X
V y (z) = −w y d z + −P y i
i =1
0

As there were no transverse distributed loads in this case, the integral


term does not come into play in this example. But we will see how it does
come into play for later examples.

Bending Moment Diagram

The two relationships for the internal bending moment diagram differ in
nature from those for the normal force and shear force diagrams. For this
diagram we have:

d Mx
= Vy
dz
∆M x = −M 0

Here we see that the slope of the internal bending moment distribution
is equal to the internal shear force distribution, while externally applied mo-
ment couples produce discontinuous steps in the bending moment diagram.
This can be observed graphically in the figure below:
1.5 Internal Force (NVM) Relationships 21

P
A θ
C
Az z
B
Cy
Ay a b

y Pb L
Pb sin θ
Area = a ⋅ sin θ L
L
Vy
Pa
− sin θ
L
Pab
sin θ
L

+
Mx
Pb Pa
slope = sin θ slope = − sin θ
L L

Figure 1.32 Construction of an M -diagram graphically from the FBD.


In this particular example, it is useful to look at the relations from the
alternative perspective of a summation of infinitesimal elements:

Zz n
X
M x (z) = Vy d z + −M 0i
i =1
0

where the integral portion of the equation is simply the area under the V -
diagram up to point z. This is useful for quickly determining the magnitude
of key values in the M -diagram, such as the magnitude of the maximum
internal bending moment that occurs at point B along the beam. Here,
we can see that there are no applied moment couples along the beam, so
the internal moment at B is simply equal to the area under the shear force
diagram from A to B .

You might also like