You are on page 1of 166

Lecture 6.

1
Reaction Influence Lines

This lecture explores the use of a qualitative approach to constructing reaction influence lines
for statically determinate beams.

Suppose we are asked to design Column BG in the bridge shown in Figure 6.1.

Figure 6.1: A bridge structure

To correctly design the column, we need to determine the maximum axial force that it must be
able to carry. This means we need to calculate the maximum reaction force at the support at
Point B.

It is important to note that the beam’s support reactions are not constant; they change as a
vehicular load moves across the bridge. We can determine the maximum reaction force at Point
B due to a moving load by drawing a diagram called an influence line. An influence line will
show how the reaction force changes as a unit load moves from Point A to Point F. The influence
line for the reaction force at Point B is shown in Figure 6.2.

Figure 6.2: The influence line for the reaction force at Point B

Chapter 6: Influence Lines Page 6.2


Before we discuss how to construct an influence line, let’s see how they are meant to be
interpreted.

The influence line in Figure 6.2 shows the values for the reaction force at Point B as a unit load
moves across the bridge. To continue our earlier example, see Figure 6.3. When a unit load is at
Point A in the bridge structure, the diagram in the figure indicates that the reaction force at
Point B is zero.

Figure 6.3: The influence line value for the reaction force at Point B when the unit load is at Point A

Following along the influence line, we can see that when the unit load reaches Point B, the
reaction force at Point B becomes 1. And when the unit load is located at Point C, the influence
line indicates that the reaction force at Point B is 2 (see Figure 6.4).

Figure 6.4: The influence line value for the reaction force at Point B when the unit load is at Point C

Furthermore, the influence line indicates that the reaction force at Point B becomes negative 4
when the unit load reaches Point E (see Figure 6.5).

Figure 6.5: The influence line value for the reaction force at Point B when the unit load is at Point E

Chapter 6: Influence Lines Page 6.3


Finally, when the unit load reaches Point F, the influence line indicates that the reaction force at
Point B is zero (see Figure 6.6).

Figure 6.6: The influence line value for the reaction force at Point B when the unit load is at Point F

A visual inspection of the influence line reveals that the absolute maximum reaction force at
Point B occurs when the load is at Point E. That is, when the load reaches Point E, the reaction
force at Point B reaches its maximum negative value of 4.

Since the influence line is drawn for a moving unit load, we must multiply the values given by
the diagram by the actual load magnitude in order to obtain the correct reaction values. For
example, if a moving vehicle exerts a force of 5 kN on the bridge, then the maximum negative
reaction force at Point B can be written as (−4)(5 𝑘𝑁) = −20 𝑘𝑁.

The maximum positive reaction force can be calculated in a similar manner. The influence line
shows a maximum positive height of 2 (at Point C). Therefore, the maximum positive reaction
force at Point B can be written as (2)(5 𝑘𝑁) = 10 𝑘𝑁.

To summarize, the maximum negative support reaction at Point B develops when a vehicle is
located at the hinge at Point E. The maximum positive support reaction develops when the
vehicle is at the hinge at Point C. Therefore, Column BG must be designed for a maximum
compressive force of 10 kN and a maximum tensile force of 20 kN, as shown in Figure 6.7.

Figure 6.7: The maximum tensile and compressive forces in Column BG

Chapter 6: Influence Lines Page 6.4


We are now ready to discuss the construction of reaction influence lines for statically
determinate beams.

If a beam has no internal hinges and is a cantilever or simply supported beam, the construction
of a reaction influence line is rather straightforward. Consider the cantilever beam shown below
in Figure 6.8.

Figure 6.8: A cantilever beam subjected to a moving load

If we replace the vehicle with a unit load, we can easily see that regardless of the location of the
load, the reaction force at Point A equals 1 (see Figure 6.9).

Figure 6.9: The reaction force at the fixed support of a cantilever beam due to a moving unit load

Since the reaction force at Point A remains constant while the load travels across the beam, we
can draw the rectangular area shown in Figure 6.10. This diagram is called the reaction
influence line for the support at Point A. The rectangular area is a representation of the reaction
force magnitude as a function of the load position.

Figure 6.10: The reaction influence line for a cantilever beam

The diagram in Figure 6.10 can be viewed as a straight horizontal line. Conceptually, we can
visualize this horizontal line as Segment AB being pushed up by one unit, as depicted in Figure
6.11.

Figure 6.11: The qualitative construction of the reaction influence line for a cantilever beam

Chapter 6: Influence Lines Page 6.5


This idea of drawing the displaced shape of the beam after it is pushed up by one unit at the
support is a convenient qualitative approach for drawing reaction influence lines.

Let’s see how this idea works for a simply supported beam. Suppose we wish to construct the
influence line for the reaction force at Point B for the beam shown in Figure 6.12.

Figure 6.12: A simply supported beam

To draw the influence line, we are going to push the beam up by one unit at Point B and draw its
displaced shape, as shown in Figure 6.13.

Figure 6.13: The reaction influence line for the right support in a simply supported beam

The influence line shown in Figure 6.13 indicates that the reaction force at Point B reaches its
maximum value when the unit load is at Point B.

Let’s compare the two influence lines that we have constructed so far. The reaction influence line
for a cantilever beam and the reaction influence line for the right reaction force of a simply
supported beam are shown in Figure 6.14.

Figure 6.14: Reaction influence lines for cantilever and simply supported beams

If both diagrams were drawn by pushing the support point upward, why does one diagram form
a horizontal line while the other forms an inclined line?

Note that the right end of the cantilever beam is free. Therefore, when we push the beam’s left
end up, its (free) right end moves up as well. In the case of the simply supported beam, however,
the roller at Point A is not free to move up; it can only rotate. Therefore, when the end at Point B
is pushed up by one unit, the beam must rotate counterclockwise to accommodate the
movement, resulting in an inclined line.

Let’s consider another example. See the statically determinate beam shown in Figure 6.15.

Figure 6.15: A beam with an internal hinge

Chapter 6: Influence Lines Page 6.6


The influence line for the reaction force at Point A can be constructed by pushing the beam up
by one unit at the end at Point A. Note that Segment BD, since it rests on a roller (at Point C)
and a pin (at Point D), cannot move up or down or rotate. Therefore, when the end at Point A is
pushed up, Segment AB turns clockwise, given that the hinge at Point B allows a relative
rotation between Segment AB and Segment BD. This forms the displaced shape shown in Figure
6.16.

Figure 6.16: The displaced shape of a beam with an internal hinge, when it is pushed up at its left
support

Therefore, the influence line for the reaction force at Point A can be drawn in the manner shown
in Figure 6.17.

Figure 6.17: The influence line for the left reaction force for a beam with an internal hinge

Now, let’s draw the influence line for the reaction force at the support at Point C. To do so, we
need to draw the displaced shape of the beam when Point C is pushed up by one unit. Segment
BD is pin-connected at Point D and has a hinge at Point B. Therefore, if Point C is pushed up,
the beam is going to displace as shown in Figure 6.18.

Figure 6.18: The displaced shape of a beam with an internal hinge, when it is pushed up at an interior
support

So, the influence line for the reaction force at Point C can be drawn as shown in Figure 6.19.

Figure 6.19: The influence line for the reaction force at an interior support for a beam with an internal
hinge

What does the influence line for the reaction force at Point D look like? To answer this question,
we need to determine the displaced shape of the beam when Point D is pushed up by one unit.

Chapter 6: Influence Lines Page 6.7


Since Point C cannot move up or down and can only rotate, if we push up Point D, Point B must
move down. This causes Segment BD to rotate counterclockwise (see Figure 6.20). Note that
since the hinge at Point B has moved down, Segment AB has rotated clockwise to accommodate
the downward displacement of its right end.

Figure 6.20: The displaced shape of a beam with an internal hinge, when it is pushed up at its right
support

Consequently, the influence line for the reaction force at Point D can be drawn as shown in Figure
6.21.

Figure 6.21: The influence line for the reaction force at the right support for a beam with an internal
hinge

Let’s consider the statically determinate beam shown in Figure 6.22.

Figure 6.22: A statically determinate beam with an internal hinge and a fixed support

Suppose we wish to draw the influence line for the reaction force at the support at Point A. As
we’ve learned, this can be done by pushing the end at Point A up by one unit and drawing the
resulting displaced shape of the beam (see Figure 6.23).

Figure 6.23: The influence line for the reaction force at the left end of a beam with an internal hinge and
a fixed support

Here, Segment CB cannot move or rotate because of the fixed support at Point B.

It is important to note that when drawing influence lines for statically determinate beams, we
treat the beam segments as straight bars. These bars are unable to bend. You might be tempted
to draw the influence line for the reaction at Point A in the manner shown in Figure 6.24.

Figure 6.24: An incorrectly drawn influence line for a statically determinate beam

Chapter 6: Influence Lines Page 6.8


However, the diagram is not correct as it is drawn in Figure 6.24 since Segment CB is not drawn
using a straight line. Influence lines for statically determinate beams always consist of straight-
line segments.

Let’s draw the influence line for the reaction force at Point B.

To do so, we need to push up the end at Point B by one unit and draw the resulting displaced
shape of the beam. When Point B is pushed up, Point C is going to move up as well, unless
Segment AC restrains its movement. In this case, since Point A is free to rotate, Segment AC can
rotate counterclockwise to accommodate an upward movement of the hinge at Point C.
Therefore, the reaction influence line for Point B can be drawn as shown in Figure 6.25.

Figure 6.25: The influence line for the reaction force at the fixed end of a beam with an internal hinge
and a fixed support

We will discuss shear and moment influence lines in the next lectures.

Chapter 6: Influence Lines Page 6.9


Exercise Problems:

Draw the reaction influence lines for each beam.

1 a) Reaction at A
b) Reaction at B

2 a) Reaction at A
b) Reaction at B

a) Reaction at A
3 b) Reaction at B
c) Reaction at C

a) Reaction at A
4 b) Reaction at B
c) Reaction at C
d) Reaction at D

Chapter 6: Influence Lines Page 6.10


Structural Analysis I Reaction Influence Line Exercise Problem Solutions

Problem 1: Draw the influence line for the reaction force at A and B.

A B

Solution:

A B

Influence line for the reaction force at A

Influence line for the reaction force at B

Dr. Structure P a g e |1
Structural Analysis I Reaction Influence Line Exercise Problem Solutions

Problem 2: Draw the influence line for the reaction force at A and B.

A B

Solution:

A B

Influence line for the reaction force at A

Influence line for the reaction force at B

Dr. Structure P a g e |2
Structural Analysis I Reaction Influence Line Exercise Problem Solutions

Problem 3: Draw the influence line for the reaction force at A, B, and C.

B C
A

Solution:

B C
A

Influence line for the reaction force at A

Influence line for the reaction force at B

Influence line for the reaction force at C

Dr. Structure P a g e |3
Structural Analysis I Reaction Influence Line Exercise Problem Solutions

Problem 4: Draw the influence line for the reaction force at A, B, C, and D.

A B C D

Solution:

A B C D

Influence line for the reaction force at A

Influence line for the reaction force at B

Influence line for the reaction force at C

Influence line for the reaction force at D

Dr. Structure P a g e |4
Lecture 6.2
Shear Influence Lines

This lecture presents a qualitative approach to constructing shear influence lines for statically
determinate beams.

Figure 6.26: A statically determinate beam with internal hinges

Consider the bridge shown in Figure 6.26. It is subjected to a vehicular (moving) load. Suppose
we are asked to design the hinge at Point C. To do so requires a determination of the maximum
shear force produced by the moving load at Point C.

The maximum positive and negative shear force values at Point C can be determined with the
aid of a shear influence line. Similar to a reaction influence line, a shear influence line is a
diagram that shows the value of a shear force at a specific point as a function of the position of a
moving unit load. The influence line for the shear force at Point C in the bridge due to the
vehicular load is shown in Figure 6.27.

Figure 6.27: Shear influence line for Point C

Chapter 6: Influence Lines Page 6.11


Before we learn how to draw this diagram, let’s first understand how to interpret it. As
mentioned, the influence line shows the shear values at Point C as a unit load moves across the
bridge. For example, when the unit load is to the left of Point C, the line diagram shows that the
shear at Point C is zero (see Figure 6.28).

Figure 6.28: Reading the shear influence line when the load is to the left of Point C

We can also see that when the unit load crosses over Point C, the shear at Point C jumps from
zero to one (see Figure 6.29).

Figure 6.29: Reading the shear influence line when the load is at and just to the right of Point C

Following along the influence line in Figure 6.29, the shear at Point C then continues to decrease
linearly as the load continues to move to the right toward Point D. When the load reaches Point
D, the shear at Point C becomes zero once again.

The shear continues to decrease linearly as the load continues to move toward Point E. As shown
in Figure 6.30, when the load reaches Point E, the shear at Point C reaches its maximum
negative value of 3/2.

Chapter 6: Influence Lines Page 6.12


Figure 6.30: Reading the shear influence line when the load is at Point E

Once the moving load crosses Point E, the shear force at Point C begins to increase from
negative 3/2 toward zero. It reaches zero when the unit load reaches Point F.

Keep in mind that since the influence line is drawn for a unit load, we must multiply the values
given by the diagram by the actual magnitude of the load in order to obtain the correct shear
value. For example, if the moving vehicle exerts a force of 6 kN on the bridge, then the maximum
negative shear force at Point C (occurring when the unit load is at Point E) equals (6 kN)(3/2),
or 9 kN. Therefore, the hinge at Point C should be designed to withstand a force of 9 kN. See
Figure 6.31.

Figure 6.31: The maximum negative shear force at the hinge at Point C due to a vehicular load

Now that we know how to read a shear influence line diagram, let’s go back and learn how to
draw one. Consider the simply supported beam shown in Figure 6.32. Suppose we wish to draw
the influence line for the shear force at Point C. That is, we want to show the shear values at
Point C as a unit load moves across the beam.

Chapter 6: Influence Lines Page 6.13


Figure 6.32: A simply supported beam

If we place the unit load at several locations on the beam and calculate the shear force at Point C
for each load position, we can plot our calculated shear values as a function of the load position.
By doing so, we create a graph similar to the one shown in Figure 6.33.

Figure 6.33: The plot of shear values for Point C

If we connect these points with straight lines, we get the shear influence line for Point C (see
Figure 6.34).

Figure 6.34: The shear influence line for Point C

Figure 6.34 tells us that the shear at Point C decreases linearly from zero to a negative value as
the unit load moves from Point A to Point C. At Point C, the shear value jumps from negative to
positive. This means that when the unit load is just to the left of Point C, the shear at Point C is
negative. When the unit load moves just to the right of Point C, the shear at Point C becomes
positive. The shear at Point C then decreases linearly again toward zero as the load moves from
Point C to Point B.

The shear diagram in Figure 6.34 consists of two line segments separated at Point C. Visually,
the diagram looks as if the beam were cut at Point C, with the left side pushed down and the

Chapter 6: Influence Lines Page 6.14


right side pushed up. In this scenario, Point A and Point B rotate to accommodate the
movement. The idea of pushing the beam segments up or down and drawing the resulting shape
is a convenient way to draw shear influence lines qualitatively.

To better apply this qualitative approach for drawing shear influence lines, we can imagine a
pair of vertical rollers placed at Point C, permitting vertical movement at that point (see
Figure 6.35).

Figure 6.35: Imaginary vertical rollers at Point C

Then, picture applying a positive shear force to these rollers. A positive shear force in beams is a
downward force applied to the left roller and an upward force applied to the right roller, as
shown in Figure 6.36.

Figure 6.36: A positive shear force placed at Point C

These forces are imagined to have the ability to move the two rollers in opposite directions. The
downward shear force causes a downward movement of the left roller, and the upward shear
force causes an upward movement of the right roller, as depicted in Figure 6.37.

Figure 6.37: Beam segments’ rotation due to an imaginary pair of shear forces

Next, we draw the centerlines of the two beam segments and connect them at Point C. Doing so
renders a line, which we refer to as the influence line for shear at Point C (see
Figure 6.38).

Chapter 6: Influence Lines Page 6.15


Figure 6.38: Shear influence line for Point C

Now let’s draw the influence line for shear at Point C for the beam shown in Figure 6.39, which
has an overhang.

Figure 6.39: A statically determinate beam with an overhang

First, we place a pair of imaginary vertical rollers at Point C, as shown in Figure 6.40. Then, we
apply a positive shear force to the rollers.

Figure 6.40: A pair of imaginary vertical rollers placed at Point C

The rollers divide the beam into two rigid segments, which we will call bars. In this example, the
rollers create Bar AC and Bar CE. How will these bars deform?

To determine the displaced shape of the bars, it is important to keep in mind that they can move
up or down or rotate, if permitted by their boundary conditions, but they cannot bend. As such,
any deformation resembling that shown in Figure 6.41 is not permissible.

Figure 6.41: An invalid deformation of a beam segment due to a shear force

Chapter 6: Influence Lines Page 6.16


To determine the displaced shape of Bar AC, we observe that Point C is being pushed down by a
shear force. Since the pin at Point A permits the bar to rotate, Bar AC will have a clockwise
rotation, as shown in Figure 6.42.

Figure 6.42: Displacement of Bar AC due to the applied shear force at Point C

Under the same shear force, how will Bar CE displace? Here, the left end of the bar is being
pushed up by the applied shear force at Point C. Given that the roller at Point B permits rotation
(a pin or a roller support can rotate) and the end at Point E is free, it follows that the upward
movement of Point C will make the entire bar rotate in a clockwise direction around Point B (see
Figure 6.43).

Figure 6.43: Displacement of Bar CE due to the applied shear force

By drawing the centerlines of the two beam segments and connecting them together at Point C,
we get the shear influence line shown in Figure 6.44.

Figure 6.44: Influence line for the shear at Point C

Now let’s draw the influence line for shear at Point D for the same beam that was shown in
Figure 6.39. First, we place an imaginary pair of vertical rollers at Point D. Then we subject the
rollers to a positive shear force, as shown in Figure 6.45.

Chapter 6: Influence Lines Page 6.17


Figure 6.45: A pair of imaginary vertical rollers placed at Point D

Now we draw the deformed shape of the beam. The rollers divide the beam into two bars: Bar
AD and Bar DE. Under this shear force, Bar DE moves up vertically, as shown in Figure 6.46.
But why does the entire length of Bar DE move up? Since Bar DE has no support points that
enable rotation, both ends of the segment are free to move up vertically.

Figure 6.46: Displacement of Bar DE due to the applied shear force at Point D

Bar AD, however, does not have any deformation at all. While it is true that Point D is being
pushed down, Point D would correspondingly move down if—and only if—Bar AD was allowed
to rotate around Point B. But for Bar AD to rotate around Point B, Point A must be able to move
up. Since Point A is pinned, it is not allowed to vertically displace.

As such, Bar AD cannot move under the downward shear force at Point D. Therefore, the entire
influence line for shear at Point D has a nonzero area only to the right of Point D, as shown in
Figure 6.47.

Figure 6.47: Influence line for the shear at Point D

Next, let’s consider the cantilever beam shown in Figure 6.48.

Figure 6.48: A cantilever beam

Chapter 6: Influence Lines Page 6.18


Let’s say we wish to draw the influence line for the shear force at Point C. To do so we first place
a pair of imaginary vertical rollers at Point C, and then we subject the rollers to a positive shear
force, as shown in Figure 6.49.

Figure 6.49: A pair of imaginary vertical rollers placed at Point C

Now we draw the displaced shape that results. The beam is divided into two bars: Bar AC and
Bar CB. Bar CB moves up since both points C and B are free to move vertically. Bar AC, however,
remains undeformed since Point A is fixed and cannot move or rotate (see Figure 6.50).

Figure 6.50: Displacement of Bar CB due to a shear force at Point C

The complete influence line for the shear at Point C is shown in Figure 6.51.

Figure 6.51: Influence line for the shear at Point C

Now, let’s consider beams with one or more hinges. Figure 6.52 depicts a statically determinate
beam with an internal hinge at Point D.

Figure 6.52: A beam with an internal hinge at Point D

Chapter 6: Influence Lines Page 6.19


Suppose we wish to draw the shear influence line for Point E. We start by placing a pair of
vertical rollers at Point E and then subject them to a positive shear force, as shown in
Figure 6.53.

Figure 6.53: A beam with an internal hinge subjected to a shear force at Point E

The hinge and the rollers divide the beam into three bars: Bar AD, Bar DE, and Bar EC. For Bar
EC, the upward force pushes Point E up as the entire bar rotates around Point C, as shown in
Figure 6.54.

Figure 6.54: Displacement of Bar EC due to a shear force at Point E

For Bar DE, we have a force pushing down at Point E. This means that the beam has to rotate
clockwise at Point B, which forces Point D to move up. For Point D to have an upward
displacement, Bar AD must then rotate counterclockwise, as shown in Figure 6.55.

Figure 6.55: Displacement of bars AD and DE due to a shear force at Point E

Finally, we draw the centerlines of each bar and connect them at Point E, which gives us the
influence line for shear at Point E (see Figure 6.56).

Figure 6.56: Influence line for the shear at Point E

Chapter 6: Influence Lines Page 6.20


Now, let’s draw the influence line for the shear at the hinge at Point D for the same beam that
was shown in Figure 6.52. We start by placing an imaginary pair of vertical rollers at Point D,
and then we apply a unit shear force to the rollers, as shown in Figure 6.57.

Figure 6.57: A pair of imaginary vertical rollers placed at Point D

The downward shear force at Point D wants to rotate Bar AD in a clockwise direction. Since the
pin at Point A permits rotation, the bar does indeed rotate, accommodating the downward
displacement. Conversely, the upward shear force on the left end of Bar DC wants to lift the end
at Point D. However, for Point D to move up, Bar DC needs to be able to rotate around Point B.
For that to happen, Point C needs to move down, but the roller at Point C prevents any vertical
displacement. Therefore, no rotation at Point B nor upward movement at Point D can occur (see
Figure 6.58).

Figure 6.58: Displacement of Bar AD due to a shear force at Point D

Therefore, the shear influence line for Point D can be drawn as shown in Figure 6.59.

Figure 6.59: Influence line for the shear at Point D

Now, let’s look at an example involving two hinges (see Figure 6.60).

Figure 6.60: A beam with two internal hinges

Chapter 6: Influence Lines Page 6.21


For this beam, what does the influence line for shear at Point E look like? To find out, let’s start
by placing a pair of imaginary vertical rollers at Point E and subjecting them to a positive shear
force, as shown in Figure 6.61.

Figure 6.61: A pair of imaginary rollers placed at Point E

Now we can draw the resulting deformed shape of the beam. Here the beam is divided into four
bars: Bar AF, Bar FE, Bar EG, and Bar GD. Due to the upward shear force at Point E, Bar EG
wants to rotate clockwise. However, it can do so only if Point G (one of the internal hinges) is
permitted to move down in Bar GD. Since Point D has a roller support, Bar GD can rotate. The
deformed shape of bars EG and GD are as shown in Figure 6.62.

Figure 6.62: Displacement of bars EG and GD due to a shear force at Point E

Similarly, Bar FE wants to rotate clockwise, since the end at Point E is being pushed down.
However, Bar FE can rotate only if the hinge at Point F is permitted to move up in response.
Since Point A has a pin support permitting rotation, the hinge at Point F can move up, as
depicted in Figure 6.63.

Figure 6.63: Displacement of bars AF and FE due to a shear force at Point E

The resulting influence line for shear at Point E is as shown in Figure 6.64.

Chapter 6: Influence Lines Page 6.22


Figure 6.64: Influence line for the shear at Point E

Next, let’s talk about determining shear magnitudes in influence lines. Consider the simply
supported beam shown in Figure 6.65.

Figure 6.65: A simply supported beam having a length of 4L

Suppose we wish to draw the influence line for the shear at Point C. We start by placing an
imaginary pair of vertical rollers at Point C and then applying a positive shear force to the
rollers, as shown in Figure 6.66.

Figure 6.66: A pair of imaginary vertical rollers placed at Point C

The rollers divide the beam into two bars (Bar AC and Bar CB), which then displace, as shown in
Figure 6.67.

Figure 6.67: Displacement of bars AC and CB due to a shear force at Point C

The resulting influence line is as shown in Figure 6.68.

Figure 6.68: Influence line for the shear at Point C

Let’s refer to the vertical distances from the x-axis to the bottom and top of the influence line at
Point C as ℎ1 and ℎ2 , respectively (see Figure 6.69).

Chapter 6: Influence Lines Page 6.23


Figure 6.69: The heights of the influence line at Point C

Since the influence line is drawn for a unit load, we know that the total vertical distance at Point
C representing the change in shear value is 1. We can express this as ℎ1 + ℎ2 = 1.

Note how the influence line forms two similar triangles (see colored triangles in Figure 6.70).

Figure 6.70: Influence line forming two similar triangles

We can use similar triangle side ratios to find ℎ1 and ℎ2 . That is, 𝐿/ℎ1 = 3𝐿/ℎ2 , or ℎ2 = 3ℎ1 .

And since ℎ1 + ℎ2 = 1, we can solve for ℎ1 and ℎ2 as follows: ℎ1 = 1/4 and ℎ2 = 3/4. With these
height values, the new, accurate shape of the influence line is shown in Figure 6.71.

Figure 6.71: Influence line for the shear at Point C

As another example, let’s consider the beam shown in Figure 6.72.

Figure 6.72: A beam with an internal hinge

Suppose we wish to draw the influence line for the shear at Point D. As usual, we place a pair of
vertical rollers at Point D, subject the rollers to a positive shear force, and then draw the
displaced shape of the beam as shown in Figure 6.73.

Chapter 6: Influence Lines Page 6.24


Figure 6.73: Influence line for the shear at Point D

The part of the influence line between Point B and Point C can be viewed as two right triangles.
Let’s label the two heights at Point D as ℎ1 and ℎ2 , as shown in Figure 6.74.

Figure 6.74: Influence line forming two similar right triangles

Using the basic geometric relationships between the two triangles, we can write the following:
ℎ1 /2 = ℎ2 /4. Further, since we know that ℎ1 + ℎ2 = 1, the two heights can be determined as
follows: ℎ1 = 1/3 and ℎ2 = 2/3. The influence line can then be drawn to scale, as shown in
Figure 6.75.

Figure 6.75: Influence line for the shear at Point D

Chapter 6: Influence Lines Page 6.25


We can determine the height of the diagram at the internal hinge using similar triangles as well.
Note the colored triangles in Figure 6.76.

Figure 6.76: Similar triangles for the center part of an influence line

The height-to-base ratio of the two triangles give us the equation 4/ℎ = 2/(1/3). Therefore, ℎ =
2/3.

Now, let’s go back and consider the influence line shown in Figure 6.51. What is the height of the
diagram at Point C?

We know that the overall height between the rollers is always one. In this case, because the
influence line has only one height located on the positive side of the x-axis, that height must be
equal to one (see Figure 6.77).

Figure 6.77: Influence line for shear in a cantilever beam

Let’s wrap up this lecture by drawing the influence line for the shear at the hinge in the beam
shown in Figure 6.72. We start by placing a pair of vertical rollers at the hinge (Point E), and
then apply a positive shear force to the rollers, as shown in
Figure 6.78.

Figure 6.78: A pair of vertical rollers and a shear force placed at the hinge at Point E

Chapter 6: Influence Lines Page 6.26


Under the unit shear force, Bar AE turns clockwise, since the downward shear force at Point E
pushes the right end of the beam down while the left end of the beam rotates around Point A. On
the other side of the hinge, the upward shear force at Point E wants to cause a clockwise rotation
of Bar EC. For this rotation to happen, Bar EC needs to rotate around Point B. The roller at
Point C, however, prevents such rotation. Consequently, Bar EC remains undeformed, as
depicted in Figure 6.79.

Figure 6.79: Influence line for the shear at Point E

As such, the influence line has a nonzero triangular shape only at the left end of the beam (from
Point A to Point E). The triangle has a height of one at Point E where the hinge is located. The
remaining part of the diagram is flat along the x-axis, signifying that the shear at Point E
remains zero once the unit load moves to the right of the hinge.

Chapter 6: Influence Lines Page 6.27


Structural Analysis I Shear Influence Line Exercise Problem Solutions

Problem 1: Draw the influence line for the shear force at hinge C.

2m
B

4m

3m

Solution:

Dr. Structure P a g e |1
Structural Analysis I Shear Influence Line Exercise Problem Solutions

Problem 2: Draw the influence line for the shear force at points F, G, H, I, J, and K.

F G H I J K
A B C D E

Solution:

Dr. Structure P a g e |2
Structural Analysis I Shear Influence Line Exercise Problem Solutions

Dr. Structure P a g e |3
Structural Analysis I Shear Influence Line Exercise Problem Solutions

Problem 3: Draw the influence line for the shear force at points A and D.

A B C D
2m 2m 3m 2m 1m

Solution:

Dr. Structure P a g e |4
Lecture 6.3
Moment Influence Lines

This lecture explores the use of a qualitative approach to constructing moment influence lines
for statically determinate beams.

Consider the bridge shown in Figure 6.80: A bridge structure. It is designed to carry moving
loads. Suppose we are asked to determine the vehicle location(s) that result in maximum
bending moment values at Point G.

Figure 6.80: A bridge structure

To determine the desired vehicle locations, we can draw and use the moment influence line for
Point G. The influence line is shown in Figure 6.81.

Figure 6.81: The moment influence line for Point G

Before we discuss how to draw a moment influence line, let’s first discuss how a moment
influence line diagram is meant to be interpreted.

Chapter 6: Influence Lines Page 6.28


The influence line in Figure 6.81 identifies the load locations that cause the bending moment at
Point G to reach its maximum values. For example, see Figure 6.82. When a unit load is at Point
A, the diagram indicates that the bending moment at Point G is zero.

Figure 6.82: Bending moment at Point G is zero when the unit load is at Point A

When the unit load reaches Point B, the bending moment at Point G has a large negative value
(see Figure 6.83).

Figure 6.83: Bending moment at Point G has a large negative value when the unit load is at Point B

When the unit load reaches Point C, the bending moment at Point G becomes zero again (see
Figure 6.84).

Figure 6.84: Bending moment at Point G becomes zero when the unit load reaches Point C

Using the moment influence line, we can also see that the bending moment at Point G reaches
its maximum positive value when the unit load reaches Point G (see Figure 6.85).

Chapter 6: Influence Lines Page 6.29


Figure 6.85: Bending moment at Point G has a large positive value when the unit load is at Point G

Then, the value of the bending moment at Point G decreases as the load moves toward Point D.
When the load reaches Point D, the moment at Point G reaches zero (see Figure 6.86).

Figure 6.86: Bending moment at Point G becomes zero when the unit load reaches Point D

The value of the bending moment at Point G continues to decrease as the unit load moves
toward Point E. When the load is at Point E, the moment at Point G reaches a large negative
value, as shown in Figure 6.87.

Figure 6.87: Bending moment at Point G reaches a large negative value when the unit load is at Point E

Then, the moment value starts moving back toward zero as the unit load moves toward Point F.
When the load reaches Point F, the moment at Point G becomes zero again (see Figure 6.88).

Chapter 6: Influence Lines Page 6.30


Figure 6.88: Bending moment at Point G becomes zero when the unit load reaches Point F

Given the influence line for the bending moment at Point G, we can easily determine the vehicle
location that causes the moment to reach its maximum positive value. As shown in
Figure 6.89, the vehicle must be located at Point G in order for the bending moment at Point G
to reach its maximum positive value.

Figure 6.89: The vehicle location for producing the maximum positive moment at Point G

So, where should the vehicle be located in order for the moment at Point G to reach its
maximum negative value? The influence line in Figure 6.89 shows two peak negative values: one
at Point B and one at Point E. This means that the maximum negative bending moment at Point
G occurs when two cars, not one, are on the bridge. The bending moment at Point G becomes
maximally negative when one vehicle is at Point B while another vehicle is at Point E, as shown
in Figure 6.90.

Figure 6.90: The vehicle locations for producing the maximum negative moment at Point G

Chapter 6: Influence Lines Page 6.31


As illustrated above, a moment influence line can facilitate the determination of moving load
patterns that result in maximum bending moments developing at a specific point in the beam.

Let’s see how such an influence line can be drawn qualitatively. Consider the simply supported
beam shown in Figure 6.91. Suppose we wish to construct the influence line for bending
moment at Point C.

Figure 6.91: A simply supported beam

The moment influence line can be constructed in three steps:

1. Place a fictitious hinge at the point of interest. In this case, we place the hinge at Point C,
as shown in Figure 6.92.

Figure 6.92: A simply supported beam with a fictitious hinge

2. Apply a positive bending moment to the hinge, as shown in Figure 6.93. A positive
bending moment manifests itself in a pair of moments: a counterclockwise moment at
the left side of the hinge, and clockwise moment at the right side of the hinge.

Figure 6.93: A simply supported beam subjected to a positive bending moment at a fictitious hinge

3. Draw the displaced shape of the beam that results from the introduction of the fictitious
hinge and the positive moment.

To draw the displaced shape of the beam, we notice that the left segment (Segment AC) wants to
turn counterclockwise because of the counterclockwise moment at Point C. And the right
segment (Segment CB) wants to turn clockwise because of the clockwise moment at Point C. For
Segment AC to turn counterclockwise, the right end of the segment must move up, as shown in
Figure 6.94.

Figure 6.94: A beam segment turning counterclockwise due to a bending moment

Chapter 6: Influence Lines Page 6.32


Similarly, for Segment CB to turn clockwise, the left end of the segment must move up, as shown
in Figure 6.95.

Figure 6.95: A beam segment turning clockwise due to a bending moment

Therefore, the entire beam displaces as shown in Figure 6.96.

Figure 6.96: The displaced shape of a simply supported beam with a fictitious hinge subjected to a
positive moment

It is important to keep in mind that a rigid beam cannot be displaced as shown in Figure 6.96.
However, the presence of the fictitious hinge at Point C permits such a displacement to occur.

The displaced shape of the beam shown in Figure 6.96 represents the moment influence line for
Point C. Keep in mind that this is not the actual displacement of the beam, as the hinge and the
placed moments at Point C are not real. They are placed at Point C as a part of the qualitative
method for drawing moment influence lines.

Let’s consider another example. A statically determinate beam with an overhang is shown in
Figure 6.97. Suppose we wish to draw the moment influence line for Point B.

Figure 6.97: A statically determinate beam with an overhang

In the same way as before, we place a fictitious hinge at Point B and subject it to a positive
moment, as shown in Figure 6.98.

Figure 6.98: A fictitious hinge subjected to a positive moment in a beam with an overhang

We then draw the resulting displaced shape of the beam. When doing so, we want to make sure
that the beam segments, which we’ve previously called bars, remain straight. That is, a bar can
move up or down as depicted in Figure 6.99, or rotate as shown in Figure 6.100. But bars cannot
bend as shown in Figure 6.101.

Chapter 6: Influence Lines Page 6.33


Figure 6.99: Permissible (up/down) displacement of a beam segment when drawing influence lines

Figure 6.100: Permissible (clockwise/counterclockwise) rotation of a beam segment when drawing


influence lines

Figure 6.101: Impermissible deflection of a beam segment when drawing influence lines

For the beam shown in Figure 6.98, the clockwise moment at the right side of the fictitious hinge
forces Bar BC to rotate clockwise, as shown in Figure 6.100. However, Bar AB cannot displace or
rotate since the pin and roller at points A and B prevent the bar’s movement. Therefore, the
displaced shape of the beam shown in Figure 6.100 constitutes the moment influence line for
Point B. Hence, we can draw the influence line as follows in Figure 6.102.

Figure 6.102: The influence line for bending moment at Point B

This influence line indicates that no bending moment develops at Point B as long as the unit
load is located to the left of Point B. When the unit load moves to the right of Point B, a negative
moment develops at Point B, which reaches its maximum value when the unit load reaches Point
C (see Figure 6.103).

Figure 6.103: The unit load position that produces the maximum moment at Point B

Let’s consider another example. The beam shown in Figure 6.104 is fixed at Point A and rests on
a roller at Point B. There is a real hinge at Point C. Suppose we wish to draw the moment
influence line for Point C.

Chapter 6: Influence Lines Page 6.34


Figure 6.104: A statically determinate beam with a fixed and a roller support

Regardless of the position of a unit load, bending moment at an internal hinge is always zero.
The following diagram in Figure 6.105 represents the influence line for the bending moment at
Point C.

Figure 6.105: The influence line for bending moment at an internal hinge

Consider the statically determinate beam shown in Figure 6.106. Let’s draw the influence line
for bending moment at Point E.

Figure 6.106: A statically determinate beam with a pin and two roller supports

We start by placing a fictitious hinge at Point E and subject it to a positive moment, as shown in
Figure 6.107.

Figure 6.107: A statically determinate beam with a fictitious hinge and a real hinge

Then, we draw the displaced shape of the beam. The counterclockwise moment at the left side of
the hinge at Point E causes Bar AE to rotate in the counterclockwise direction. Consequently,
Point E moves, forcing Bar EC to rotate clockwise around the roller at Point B. For Bar EC to
rotate clockwise, the hinge at Point C must move down, forcing Bar CD to rotate in the
counterclockwise direction. The resulting displaced shape of the beam is shown in Figure 6.108.

Figure 6.108: The displaced shape of a beam with a real hinge and a fictitious hinge subjected to a
positive moment

Chapter 6: Influence Lines Page 6.35


Therefore, the moment influence line for Point E can be drawn as follows in Figure 6.109.

Figure 6.109: The moment influence line for a mid-span point for a beam with a real hinge and a
fictitious hinge

Let’s consider another example. The statically determinate beam in Figure 6.110 has an
overhang and a real hinge in the middle segment. Suppose we wish to draw the influence line for
bending moment at Point D.

Figure 6.110: A statically determinate beam with an overhang and a real hinge

As usual, we place a fictitious hinge at Point D and subject it to a positive moment, as shown in
Figure 6.111.

Figure 6.111: A statically determinate beam with an overhang, a real hinge, and a fictitious hinge

To draw the displaced shaped of the beam, it is useful to make a mental note of the fact that the
two hinges in Figure 6.111 divide the beam into three rigid bars: Bar AC, Bar CD, and Bar DF.
The clockwise moment at Point D turns Bar CD in the clockwise direction. Consequently, the
hinge at Point C moves down, forcing Bar AC to turn clockwise about Point B (see Figure 6.112).

However, although a clockwise moment is exerted at the right side of the hinge at Point D, the
supports at points E and F restrain the movement of Bar DF. Therefore, the bar cannot move or
rotate.

Figure 6.112: The displaced shape of a beam with an overhang, a real hinge, and a fictitious hinge
subjected to a positive moment

Chapter 6: Influence Lines Page 6.36


The influence line for bending moment at Point D is shown in Figure 6.113.

Figure 6.113: The moment influence line for a mid-span point for a beam with an overhang, a real
hinge, and a fictitious hinge

The qualitative approach described herein helps identify the load patterns that produce the
critical moment values at a given point. This method, however, does not actually yield numeric
values for bending moments. Further analysis is required when the moment magnitudes are
needed.

For example, suppose we are to determine the maximum negative bending moment at Point G
due to a moving load of 2 kN on the beam shown in Figure 6.114.

Figure 6.114: A statically determinate beam subjected to a moving load of 2 kN

The moment influence line for Point G is shown in Figure 6.115.

Figure 6.115: The moment influence line for Point G

The influence line tells us that the maximum negative moment at Point G develops when the
moving load is at Point F. To calculate the magnitude of that moment, we can place a unit load
at Point F, and then analyze the beam. Figure 6.116 shows the placement of a unit load at Point
F.

Figure 6.116: The beam loaded for producing the maximum negative moment at Point G

Chapter 6: Influence Lines Page 6.37


When the load is just to the right of Point F, Segment AF carries no portion of it. In this case, the
entire load is carried by Segment FD. Therefore, if we separate the beam into two segments at
the hinge at Point F, we can draw the following diagram in Figure 6.117 for the right segment of
the beam.

Figure 6.117: A beam segment subjected to a unit load

Given that the beam shown in Figure 6.117 is statically determinate, we can easily calculate its
support reactions. These reaction forces are shown in Figure 6.118.

Figure 6.118: The support reactions for a beam segment subjected to a unit load

If we cut the beam shown in Figure 6.118 at Point G, we can write the moment equilibrium
equation as follows: 𝑀𝑔 = −(0.5)(4) = −2. The result is a negative bending moment at Point G
due to a unit load placed at Point F.

To determine the magnitude of the moment due to the applied load, we can multiply the value of
𝑀𝑔 by the magnitude of the load. That is, the maximum negative bending moment at Point G
due to the load of the vehicle is (−2)(2) = −4 kN. m.

We’ll continue our discussion on influence lines in the next lecture.

Chapter 6: Influence Lines Page 6.38


Exercise Problems:

For each beam, draw the moment influence line for the circled point. Then, calculate the
maximum positive and negative moments at each circled point for a moving load of 50 kN.

Chapter 6: Influence Lines Page 6.39


Structural Analysis I Moment Influence Line Exercise Problem Solutions

Problem 1: Draw the moment influence line for points A, B, and C. Then calculate the maximum positive
and negative moments at each point. The beam is subjected to a moving load of 50 kN.

A B C

8m 4m 4m 4m 4m

Solution:

A B C

8m 4m 4m 4m 4m

Influence line for bending moment at A

Influence line for bending moment at B

Influence line for bending moment at C

Dr. Structure P a g e |1
Structural Analysis I Moment Influence Line Exercise Problem Solutions

Since the influence line for bending moment at A does not have a positive region, no positive moment
develops at A due to the moving load.

To determine the maximum negative moment at A, we can place the moving load on the beam where
the influence line peaks, as shown below.

50 kN
A B C

8m 4m 4m 4m 4m

We then analyze the beam to determine the magnitude of maximum negative moment at A.

50 kN
V B C
8m 4m 4m 4m 4m
M

Using the free-body diagram of the left segment of the beam, we can easily calculate the maximum
negative bending moment at A. It equals (8 m)(50 kN) = 400 kN.m

To determine the maximum positive moment at B, we can place the moving load on the beam where
the influence line peaks, as shown below.

50 kN
A C

8m 4m B 4m 4m 4m

Dr. Structure P a g e |2
Structural Analysis I Moment Influence Line Exercise Problem Solutions

The free-body diagram of the middle segment of the beam (the segment between the two internal
hinges) is shown below.

50 kN
25 kN 25 kN

4m B 4m

Using the above diagram, we can easily calculate the maximum positive bending moment at B. It equals
(4 m)(25 kN) = 100 kN.m
Note that since the influence line does not have any negative region, no negative bending moment
develops at B as the applied load moves across the beam.

To determine the maximum negative moment at C, we can place the moving load on the beam where
the influence line peaks, as shown below.

50 kN
A C

8m 4m B 4m 4m 4m

We then analyze the beam to determine the magnitude of maximum negative moment at C.

50 kN
A C

8m 4m B 4m 4m 4m

Using the right free-body diagram above, we can determine the maximum negative moment at C. The
moment value equals (4 m)(50 kN) = 200 kN.m

The maximum positive moment at C is zero since the influence line has no positive region.

Dr. Structure P a g e |3
Structural Analysis I Moment Influence Line Exercise Problem Solutions

Problem 2: Draw the moment influence line for points A and B. Then calculate the maximum positive
and negative moments at each point. The beam is subjected to a moving load of 50 kN.

A B
10m 5m 5m 5m
Solution:

A B
10m 5m 5m 5m

-
-

Influence line for bending moment at A

Influence line for bending moment at B

Dr. Structure P a g e |4
Structural Analysis I Moment Influence Line Exercise Problem Solutions

To determine the maximum positive moment at A, place the load at A where the influence line peaks in
the positive region, as shown below.

50 kN
B
10m 5m A 5m 5m

-
-

The free-body diagram of the beam can be drawn as shown below.

50 kN
B
10m 5m A 5m 5m
25 kN 25 kN

Using the above diagram, the moment at A can be determined as: (5 m)(25 kN) = 125 kN.m
To determine the maximum negative moment at A, we placed the load at the left end of the beam
where the influence line peaks in the negative direction.

50 kN
B
10m 5m A 5m 5m

-
-

Dr. Structure P a g e |5
Structural Analysis I Moment Influence Line Exercise Problem Solutions

The free-body diagram for the beam is shown below.

50 kN
B
10m 5m A 5m 5m
75 kN 50 kN

Hence, the maximum negative moment at A equals: (5 m)(50 kN) = 250 kN.m

Dr. Structure P a g e |6
Structural Analysis I Moment Influence Line Exercise Problem Solutions

Problem 3: Draw the moment influence line for points A and B. Then calculate the maximum positive
and negative moments at each point. The beam is subjected to a moving load of 50 kN.

Solution:

The influence lines for the bending moments at points A and B are shown below.

To determine the maximum negative moment at point A, place the load on the beam where the
influence line peaks in the negative region, as shown below.

Dr. Structure P a g e |7
Structural Analysis I Moment Influence Line Exercise Problem Solutions

Then, draw the free-body diagram for the beam and determine the support reactions.

Using the free-body diagram of the middle segment of the beam shown above, we can calculate the
bending moment at point A. The moment equals (4 m)(50 kN) = 200 kN.m

Since the influence line for bending moment at point A does not have any positive region, the maximum
positive moment at A is zero.

To determine the maximum positive moment at B due to the applied load, we need to place the load on
the beam at a point that corresponds to the peak positive value on the influence line. According to the
influence line for the moment at point A (shown above), there are two such peak values: one at point B
and the other at the free end of the beam. That is, bending moment at B reaches its maximum positive
value under either load position shown below.

Case 1:

Case 2:

Dr. Structure P a g e |8
Structural Analysis I Moment Influence Line Exercise Problem Solutions

Therefore, we can use either load case to determine the maximum positive moment at point B. Using
Load Case 1, the following free-body diagram can be drawn.

The free-body diagram of the middle segment of the beam can be used to calculate the maximum
positive moment at point B. The moment equals (4 m)(25 kN) = 100 kN.m

Alternatively, Load Case 2 gives us the following free-body diagram for the beam.

Using the free-body diagram of the middle segment of the beam, we can calculate the maximum
positive moment at point B as (4 m)(25 kN) = 100 kN.m

As you can see, either loading case gives us the maximum positive moment at point B.

Dr. Structure P a g e |9
Structural Analysis I Moment Influence Line Exercise Problem Solutions

To determine the maximum negative moment at point B due to the applied load, place the load on the
beam at point B where the influence line peaks in the negative region.

The free-body diagram of the beam for the above loading case is shown below.

Using the free-body diagram of the middle segment of the beam, we can calculate the negative bending
moment at B. The moment equals (4 m)(25 kN) = 100 kN.m

Dr. Structure P a g e | 10
Lecture 6.4
Truss Influence Lines (Introduction)

This lecture focuses on the construction of influence lines for statically determinate trusses.

Consider the truss bridge shown in Figure 6.119. To design the structure properly, we need to
determine the maximum axial force that each member will be subjected to as a load moves
across the bridge. This can be accomplished using influence lines.

Figure 6.119: A truss bridge

Since an influence line is drawn for a unit load, we start by placing a moving unit load on the
bridge, as shown in Figure 6.120.

Figure 6.120: A truss subjected to a moving unit load

Suppose we wish to determine the maximum axial force in Member DJ. The influence line for
the force in the member is shown in Figure 6.121. From the influence line, we can see that when
the moving load is at Node D of the truss, the tensile force in the member reaches its maximum
value. For a unit load, the maximum tensile force in the member is 6⁄5 √2.

Chapter 6: Influence Lines Page 6.40


Figure 6.121: The influence line for a diagonal truss member

The regions of the influence line that lie above the x-axis correspond to the locations of the
moving load that cause a tensile force to develop in the member. The regions of the influence
line below the x-axis correspond to the load locations that cause a compressive force to develop
in the member.

According to the influence line shown in Figure 6.121, Member DJ remains in tension while the
moving load is to the left of Node D. The member stays in tension to the right of Node D until
the load nears Node E, at which point the force in the member flips from tension to
compression.

Since the influence line is drawn for a moving unit load, we need to multiply the diagram values
by the magnitude of the actual moving load in order to determine the member forces. For
example, if the moving load has a magnitude of 50 kN, as shown in Figure 6.122, we multiply 50
by 6⁄5 √2 to determine the maximum tensile force in Member DJ.

Figure 6.122: The vehicle location for inducing the maximum axial force in the truss member

Maximum Tensile Force in Member DJ = (50 kN)(6⁄5 √2) = 60√2 kN [6.1]

Chapter 6: Influence Lines Page 6.41


To draw an influence line for a bridge truss, we first need to establish the path of travel for the
load. For example, for the bridge shown in Figure 6.123, vehicles travel along the bottom chord
of the structure. In contrast, the path of travel for the bridge in Figure 6.124 is along the top
chord of the structure.

Figure 6.123: A truss bridge supporting vehicular movement along its bottom chord

Figure 6.124: A truss bridge supporting vehicular movement along its top chord

This distinction is important. To draw a member influence line, we need to be able to identify
the truss nodes that will be subjected to the moving load. And to draw the influence line for a
particular member, we move a unit load from truss node to truss node along the path of travel.
We then calculate the force in the member for each load position. This will give us a table of
member forces, which when graphed, yield the influence line for the member. Let’s illustrate
this process using an example.

Consider the bridge shown previously in Figure 6.119. Suppose we wish to draw the influence
line for Member DJ. Six nodes lie on the path of load travel. When the load is at Node A or Node
F, however, the truss members are not subjected to any axial force; instead, the entire applied
load is carried by the reaction force at the support at that node. Therefore, the focus of our
analysis is the four interior nodes that lie along the path of travel. More specifically, we need to
determine the force in Member DJ when the unit load is at Node B, Node C, Node D, and Node
E. We’ll go through each of these instances below.

Unit Load at Node B: Figure 6.125 shows the free-body diagram of the truss when the unit load
is at Node B.

Figure 6.125: A truss subjected to a unit load at its first interior node along the bottom chord

Chapter 6: Influence Lines Page 6.42


We can determine the support reaction at Node F by writing the moment equilibrium equation
about Node A, as shown.

∑ 𝑀 = 0 ⇒ (5𝐿)(𝐹𝑦 ) − (1)(𝐿) = 0 [6.2]


𝐴

Solving Equation [6.2] for the unknown force, we get 𝐹𝑦 = 1⁄5.

Then, we can use the method of sections to determine the force in Member DJ. We will cut the
truss as shown in Figure 6.126.

Figure 6.126: A vertical cut in a truss subjected to a unit load at its first interior node along the bottom
chord

We can then draw the free-body diagram of the right segment of the structure (see Figure 6.127).

Figure 6.127: The free-body diagram of the right segment of a truss subjected to a unit load at its first
interior node along the bottom chord

Since we are only interested in determining 𝐹𝑑𝑗 , we can write and solve the following
equilibrium equation for the unknown force:

1 √2
∑ 𝐹𝑦 = − 𝐹 = 0 [6.3]
5 2 𝑑𝑗
Or,

2
𝐹𝑑𝑗 = [6.4]
5√2

Unit Load at Node C: Figure 6.128 shows the free-body diagram of the truss when the unit load
is at Node C.

Chapter 6: Influence Lines Page 6.43


Figure 6.128: A truss subjected to a unit load at its second interior node along the bottom chord

We can determine 𝐹𝑦 by writing the moment equilibrium equation about Node A, as shown.

(5𝐿)(𝐹𝑦 ) − (2)(𝐿) = 0 ⇒ 𝐹𝑦 = 2⁄5 [6.5]

Then, the free-body diagram of the right segment of the structure can be drawn as follows in
Figure 6.129:

Figure 6.129: The free-body diagram of the right segment of a truss subjected to a unit load at its second
interior node along the bottom chord

To determine 𝐹𝑑𝑗 , we can write and solve the following equilibrium equation:

2 √2
∑ 𝐹𝑦 = − 𝐹 = 0 [6.6]
5 2 𝑑𝑗

Or,

4
𝐹𝑑𝑗 = [6.7]
5√2

Unit Load at Node D: Figure 6.130 shows the free-body diagram of the right segment of the
truss when the unit load is at Node D.

Chapter 6: Influence Lines Page 6.44


Figure 6.130: The free-body diagram of the right segment of a truss subjected to a unit load at its third
interior node along the bottom chord

To determine 𝐹𝑑𝑗 , we can write and solve the following equilibrium equation:

3 √2 6
∑ 𝐹𝑦 = − 𝐹𝑑𝑗 = 0 → 𝐹𝑑𝑗 = [6.8]
5 2 5√2

Unit Load at Node E: Figure 6.131 shows the free-body diagram of the right segment of the
truss when the unit load is at Node E.

Figure 6.131: The free-body diagram of the right segment of a truss subjected to a unit load at its fourth
interior node along the bottom chord

To determine 𝐹𝑑𝑗 , we can write and solve the following equilibrium equation:

4 √2 2
∑ 𝐹𝑦 = −1− 𝐹𝑑𝑗 = 0 → 𝐹𝑑𝑗 = − [6.9]
5 2 5√2

The results of the repeated analysis of the truss can be tabulated as shown in Table 6.1.

Chapter 6: Influence Lines Page 6.45


Unit Load Location Axial Force in Member DJ

A 0

2
B
5√2

4
C
5√2

6
D
5√2

2
E −
5√2

F 0

Table 6.1: Axial forces in a diagonal truss member due to a moving unit load

The graph of Table 6.1 is shown in Figure 6.132. The graph represents the influence line for
Member DJ.

Figure 6.132: The influence line for a diagonal member in a truss

Before we end this lecture, let’s consider another example. Suppose we wish to draw the
influence line for Member EJ in the truss shown in Figure 6.132.

A close examination of the truss reveals that when the moving unit load is at any node other
than Node E, Member EJ carries no force; that is, Member EJ acts as a zero-force member. But,
when the unit load is at Node E, the member carries an axial tensile force of one.

Chapter 6: Influence Lines Page 6.46


Of course, we can always analyze the truss repeatedly, as we did in the previous example, to
arrive at the final analysis results. But, given that we can easily identify the zero-force members
in a truss structure, we can also write the following Table 6.2 without any further analysis:

Unit Load Location Axial Force in Member EJ

A 0

B 0

C 0

D 0

E 1

F 0

Table 6.2: Axial forces in a vertical truss member due to a moving unit load

The influence line for Member EJ can be obtained by graphing the data given in Table 6.2, as
shown in Figure 6.133.

Figure 6.133: The influence line for a vertical member in a truss

We will expand on this discussion in the next lecture.

Chapter 6: Influence Lines Page 6.47


Exercise Problems:

(1) Draw the influence line for members HI and EF.

(2) Draw the influence line for Member PR, if the load moves between nodes H and L.

Chapter 6: Influence Lines Page 6.48


Structural Analysis I Truss Influence Line Exercise Problem Solutions

Problem 1: Draw the influence line for the axial force in members HI and EF.

B C D E F

5m
A
5m G 5m H 5m I 5m 5m K 5m L
J
Solution:

Cut the truss through member HI and draw the free-body diagrams for the left part of the structure
when a unit load is placed on the truss along its bottom chord. Two free-body diagrams can be drawn,
as shown below.
B C Fcd B C F cd
5m Fci 5m Fci
A Fhi A Fhi
H 5m G 5m
Ay 1
H
Ay
x
When the unit load is to the left of point H, the left diagram is applicable. When the unit load moves to
the right of H, the right diagram applies. We can determine Fhi in terms of A y for each diagram. For the
left diagram, we get:

x
 M = 5Fhi − 10Ay + (1)(10 − x) = 0  Fhi = 2A y − 2 +
5
0  x  10
@C

For the right diagram, we get:

 M = 5Fhi − 10Ay = 0  Fhi = 2Ay 10  x  30


@C

Now, place a unit load at a node along the bottom chord of the truss and calculate the resulting support
reaction at A, and consequently the corresponding Fhi . The following table shows the results of this
process.

Load Location x location Ay Fhi


A 0 1.00 0.00
G 5 0.75 0.50
H 10 0.50 1.00
I 15 0.25 0.50
J 20 0.00 0.00
K 25 -0.25 -0.50
L 30 -0.50 -1.00

Dr. Structure P a g e |1
Structural Analysis I Truss Influence Line Exercise Problem Solutions

To draw the influence line for the force in member Fhi , we can plot Fhi versus x, as shown below.

B C D E F

5m
A
5m G 5m H 5m I 5m 5m K 5m L
J
1
0.5 0.5
+

0.5
-
1

To draw the influence line for the force in member EF, we need to cut the truss through EF and draw the
free-body diagram for the right part of the structure, as sown below.

F F
Fef Fef
Fek 5m Fek 5m
F jk F jk
K 5m L K 5m L

x
1
Note that we have two distinct diagrams. The left diagram is applicable when the moving load is to the
left of point K. The right diagram applies to cases in which the load is to the right of point K. To
determine Fef as a function of the unit load location, we need to write two equations, one equation for
each free-body diagram. For the left free-body diagram, we can write:

 M = 5Fef = 0  Fef = 0 0  x  25
@K

That is, as long as the unit load is to the left of point K, EF is a zero-force member.

Using the right free-body diagram above, we can write:

x
 M = 5Fef − (1)(x − 25) = 0  Fef = − 5
5
25  x  30
@K

Dr. Structure P a g e |2
Structural Analysis I Truss Influence Line Exercise Problem Solutions

Tabulating the results, we get:

Load Location x location Fef


A 0 0.00
G 5 0.00
H 10 0.00
I 15 0.00
J 20 0.00
K 25 0.00
L 30 1.00

To draw the influence line for the force in member Fef , we can plot Fef versus x, as shown below.

B C D E F

5m
A
5m G 5m H 5m I 5m 5m K 5m L
J

1
+

Dr. Structure P a g e |3
Structural Analysis I Truss Influence Line Exercise Problem Solutions

Problem 2: Draw the influence line for the axial force in member PR if the unit load moves horizontally
between nodes H and L.

A B 5’ C 5’ D 5’ E 5’ F

5’
G
H I J K L
5’
M N

5’
O P

5’
Q R

5’
S T
5’

To determine the force in member PR, we can cut the truss through members OQ, OR, and PR. The free-
body diagram of the upper part of the structure is shown below.

A 5’ B 5’ C 5’ D 5’ E 5’ F

5’
G
H I J K L
5’
x
1
M N

5’
O P

Foq For Fpr

The following equilibrium equation gives us Fpr in terms of the position of the unit load.

x
 M = 5Fpr + (1)(5 + x) = 0  Fpr = −1 −
5
0  x  20
@O

Dr. Structure P a g e |4
Structural Analysis I Truss Influence Line Exercise Problem Solutions

If we move the unit load from joint H to joint L one joint at a time, we can generate the following table.

Load Location x location Fef


H 0 -1.00
I 5 -2.00
J 10 -3.00
K 15 -4.00
L 20 -5.00

To draw the influence line for the force in member Fpr , we can plot Fpr versus x, as shown below.

A B 5’ C 5’ D 5’ E 5’ F

5’
G
H I J K L
5’
M N

5’
O P

5’
Q R

5’
S T
5’

1
2
-
3
4
5

Dr. Structure P a g e |5
Lecture 6.5
Truss Influence Lines (Moving Load Series)

This lecture explains the use of influence lines for the analysis of truss bridges subjected to
moving load series.

Suppose we wish to design a truss bridge that provides a passage for vehicles across a canal (see
Figure 6.134).

Figure 6.134: A truss bridge across a canal

The structure is to be constructed such that the vehicle load transfers to the two side trusses via
a series of beams supporting the bridge deck (see Figure 6.135).

Figure 6.135: Transverse beams connect the two side trusses and support the bridge deck

The bridge is expected to support a maximum load of a three-axle truck similar to the one shown
in Figure 6.136.

Chapter 6: Influence Lines Page 6.49


Figure 6.136: A three-axle truck

As shown, the truck load can be viewed as a series of concentrated loads with fixed spacing
between each axle.

Our analysis will involve determining the maximum effect of the moving load series on a given
truss member. More specifically, we wish to determine the truck locations that cause the
maximum axial compressive and tensile forces in a target member. This problem can be solved
with the aid of influence lines.

Before we get into the analysis, let’s visualize how the truck load transfers to the truss joints as it
moves across the bridge. For this visualization, we can assume that the entire load of the truck is
concentrated at its center of gravity. As previously shown in Figure 6.135, the bridge deck rests
on seven beams, each attached to two truss joints; one joint is on each side of the beam.

When the truck enters the bridge and is directly above the leftmost beam, the entire load
transfers to that beam, as shown in Figure 6.137(a). The load is then distributed to the two truss
joints supporting the beam, as shown in Figure 6.137(b).

(a) (b)

Figure 6.137: Load transfer scheme from a bridge deck to the supporting truss joints

When the truck moves forward to a point between the first beam and the second beam, the load
is distributed between the two beams in proportion to the load’s distance from them. In turn,
the beam loads are transferred to their respective end truss joints, as shown in Figure 6.138 .

Chapter 6: Influence Lines Page 6.50


Figure 6.138: Proportional load distribution from a bridge deck to four truss joints

This pattern of load transfer shifts forward from joint to joint until the vehicle leaves the bridge.
We can use this moving load pattern to construct the influence line for a truss member.

Suppose we wish to determine the effect of the moving load on Member AB (see Figure 6.139).

Figure 6.139: A target truss member (Member AB) for moving load analysis

To draw the influence line for the member, we start by placing a unit load at the first inner joint
of the truss (from the left end of the bridge) and calculating the resulting support reactions. The
calculated reaction forces are shown in Figure 6.140.

Figure 6.140: Support reactions for a truss subjected to a unit load at its first interior joint

Then, we cut the truss through Member AB and draw the free-body diagram for the right
segment of the structure, as shown in Figure 6.141.

Chapter 6: Influence Lines Page 6.51


Figure 6.141: The free-body diagram of a segment of a truss subjected to a unit load at its first interior
joint

Using basic trigonometry, we can determine the sine and cosine of angle 𝛼 with the equations
𝑠𝑖𝑛(𝛼) = 3/√13 and 𝑐𝑜𝑠(𝛼) = 2/√13. Using the free-body diagram shown in Figure 6.141, we
can write the following equilibrium equation:

1
∑ 𝐹𝑦 = 𝐹𝑎𝑏 cos(𝛼) + =0 [6.10]
6

Solving Equation [6.10] for the unknown member force, we get 𝐹𝑎𝑏 = − √13/12.

Next, we move the load to the next truss node and calculate the resulting support reactions, as
shown in Figure 6.142.

Figure 6.142: The free-body diagram of a segment of a truss subjected to a unit load at its second
interior joint

We can then write the following equilibrium equation based on the free-body diagram of the
right segment of the truss shown in Figure 6.142:

Chapter 6: Influence Lines Page 6.52


1
∑ 𝐹𝑦 = 𝐹𝑎𝑏 𝑐𝑜𝑠(𝛼) + =0 [6.11]
3

Solving Equation [6.11] for the unknown member force, we get 𝐹𝑎𝑏 = − √13/6.

We continue moving the unit load from node to node, calculating the resulting force in Member
AB at each step. In the next calculation, the load is placed at the third interior node. The
resulting reaction forces and the free-body diagram of the right segment of the truss are shown
in Figure 6.143.

Figure 6.143: The free-body diagram of a segment of a truss subjected to a unit load at its third interior
joint

The resulting free-body diagram can be used to write the following equilibrium equation:

1
∑ 𝐹𝑦 = 𝐹𝑎𝑏 𝑐𝑜𝑠(𝛼) + =0 [6.12]
2
Solving Equation [6.12] for the unknown member force, we get 𝐹𝑎𝑏 = − √13/4.

Continuing this pattern, we place the load next at the fourth interior node. We calculate the
support reactions and draw the free-body diagram of the right segment of the truss, as shown in
Figure 6.144 .

Chapter 6: Influence Lines Page 6.53


Figure 6.144: The free-body diagram of a segment of a truss subjected to a unit load at its fourth
interior joint

The free-body diagram yields the following equilibrium equation:

2
∑ 𝐹𝑦 = 𝐹𝑎𝑏 cos(𝛼) − 1 + =0 [6.13]
3
Solving Equation [6.13] for the unknown member force, we get 𝐹𝑎𝑏 = √13/6.

Next, the load is placed at the fifth interior node. The calculated support reactions and the free-
body diagram of the right segment of the truss are shown in Figure 6.145.

Figure 6.145: The free-body diagram of a segment of a truss subjected to a unit load at its fifth interior
joint

Then, we can write the following equilibrium equation:

5
∑ 𝐹𝑦 = 𝐹𝑎𝑏 cos(𝛼) − 1 + =0 [6.14]
6
Solving Equation [6.14] for the unknown member force, we get 𝐹𝑎𝑏 = √13/12.

Let’s tabulate the unit load locations and the associated forces in Member AB in Table 6.3.

Chapter 6: Influence Lines Page 6.54


Unit Load Location (from left side) 𝐹𝑎𝑏

Left support 0

First interior node √13



12

Second interior node √13



6

Third interior node √13



4

Fourth interior node √13


6

Fifth interior node √13


12

Right support 0

Table 6.3: Axial force in Member AB and corresponding unit load location

By plotting a graph of the resulting forces in Member AB, we obtain the influence line for
Member AB, shown in Figure 6.146.

Figure 6.146: The influence line for a diagonal truss member in a bridge

Chapter 6: Influence Lines Page 6.55


To determine the maximum compressive force in the truss member, we can start by placing the
truck in the position shown in Figure 6.147. The units for the three concentrated loads are kilo-
newtons, and the distances between the loads are in meters.

Figure 6.147: The position of a truck for inducing a compressive force in a targeted truss member

To determine the axial force in Member AB due to the truck loads in the given positions, we
need to determine the heights of the influence line at locations x, y, and z, which are the
positions of the truck wheels (see Figure 6.147).

The front and middle wheels of the truck are located at the third and second interior nodes of
the truss, respectively. Therefore, the heights of the influence line at positions y and z are
known. They are −√13/12 and −√13/6, respectively.

To determine the height of the influence line at location x, we can use the height-to-base ratio
property of similar triangles. Since the rear wheels of the truck are 1 meter to the left of the
middle wheels, which are placed 3 meters to the right of the pin support, then the rear wheels
are located at 2 meters to the right of the support. Therefore, the triangle in Figure 6.148 can be
extracted from the influence line.

Figure 6.148: Two similar triangles as a part of an influence line

Chapter 6: Influence Lines Page 6.56


The height of the influence line at location x, denoted as ℎ in Figure 6.148, can be calculated as
shown:

√13
− 4 ℎ √13
= ⇒ ℎ= − [6.15]
9 2 18

We are now ready to determine the axial force in Member AB due to the series of concentrated
loads placed at the locations shown previously in Figure 6.147.

Figure 6.149: The locations of three concentrated loads and their associated influence line heights

Per the influence line and load positions shown in Figure 6.149, the axial (compressive) force in
Member AB is:

√13 √13 √13


𝐹𝑎𝑏 = (65)(− ) + (70)(− ) + (60)(− ) ≅ −70 𝑘𝑁 [6.16]
18 12 6

Equation [6.16] is the sum of the products of each load and the corresponding height of the
influence line under each load. The negative sign indicates that the axial force is compressive.

Suppose we are asked to determine the maximum compressive force in Member AB. Is it 70 kN?
Or does a different truck location yield a larger axial compressive force in the member?

To answer this question, we need to consider other truck locations that could potentially yield a
large compressive force in the member. Figure 6.150 shows such a position: the front wheels of a
truck are at the third interior joint of the truss, the middle wheels are at the second interior
joint, and the rear wheels are 1 meter to the left of the middle wheels.

Chapter 6: Influence Lines Page 6.57


Figure 6.150: A truck location for producing the maximum compressive force in a truss member

According to the influence line shown in Figure 6.150, we can calculate the axial force in
Member AB as follows:

−√13 −√13 −5√13


𝐹𝑎𝑏 = (60) + (70) + (65) ≅ −129 𝑘𝑁 [6.17]
4 6 36
Note that in Equation [6.17], the height of the influence line under the 65 kN load is given as
−5√13/36. This height was determined using the height-to-base ratio property of two similar
triangles, similar to what was done in the previous step (see Figure 6.148 and Equation [6.15]).

So, is the maximum compressive force in Member AB equal to 129 kN? Clearly, 129 kN is the
maximum compressive force in Member AB when the truck travels from the left end to the right
end of the bridge. But what if the truck travels in the opposite direction? Would the right-to-left
travel path yield a higher compressive force in the member? Let’s investigate.

Changing the truck’s direction of travel leads to the truck position shown in Figure 6.151.

Chapter 6: Influence Lines Page 6.58


Figure 6.151: A potential truck location on the bridge for producing the maximum compressive force in a
diagonal member

Using basic geometry, we can determine the heights of the influence line under the 60 kN and 70
kN loads. They are −5√13/36 and −2√13/9, respectively. Then, the axial force in the member
can be calculated as follows:

−√13 −2√13 −5√13


𝐹𝑎𝑏 = (65) + (70) + (60) ≅ −145 𝑘𝑁 [6.18]
4 9 36
According to Equation [6.18], the axial compressive force in Member AB for the truck position
shown in Figure 6.151 is 145 kN.

Clearly, as the truck moves forward toward the left end of the bridge, the axial force in the
member becomes smaller, since the influence line values associated with the load series
decrease in magnitude. Therefore, we can safely conclude that the largest compressive force in
Member AB is 145 kN. This force develops when the truck is moving in the direction and at the
location shown in Figure 6.151.

Using the same influence line, we can also determine the maximum tensile force in Member AB.
Note that the right region of the influence line defines a positive area. This means that when the
truck is on the right side of the bridge, Member AB carries a tensile force. For example, when the
truck assumes the position shown in Figure 6.152, we can determine the axial tensile force in the
member using the following equation:

√13 5√13 √13


𝐹𝑎𝑏 = (65) + (70) + (60) ≅ 86 𝑘𝑁 [6.19]
6 36 18

Chapter 6: Influence Lines Page 6.59


Figure 6.152: A potential truck location on the bridge for producing the maximum tensile force in a
diagonal member

And, when the truck assumes the position traveling in the opposite direction as shown in Figure
6.153, the tensile force in the member becomes:

√13 √13 √13


𝐹𝑎𝑏 = (60) + (70) + (65) = 70 𝑘𝑁 [6.20]
6 12 18

Figure 6.153: A potential truck reverse location on the bridge for producing the maximum tensile force
in a diagonal member

Since no other loading scenario exists that could produce a larger tensile force in Member AB
(compared to the two cases presented above), we can conclude our analysis by stating that the

Chapter 6: Influence Lines Page 6.60


maximum tensile force in the member is 86 kN. This force is developed when the truck is moving
in the direction and at the location shown in Figure 6.152.

In summary, when we have a moving load series acting on a truss structure, we need to
investigate all possible loading scenarios that could produce the largest axial force in the
member. This means that we need to move the load series in both directions up and down the
bridge in search of the locations that cause the maximum axial tensile and compressive forces to
develop in our target member.

Exercise Problems:

1) Determine the maximum compressive and tensile forces in members CD, GH, and DH.
The bridge is expected to support a moving vehicle with two axles, each exerting a
vertical load of 20 kN on the truss. The distance between the two axles is 3 meters.

2) Determine the maximum compressive and tensile forces in Member AB due to the
weight of the truck.

Chapter 6: Influence Lines Page 6.61


Structural Analysis I Truss Influence Line (Moving Load) Exercise Problem Solutions

Problem 1: Determine the maximum compressive and tensile force in members 𝐶𝐶𝐶𝐶, 𝐺𝐺𝐺𝐺 and
𝐷𝐷𝐷𝐷. The structure is expected to support a moving vehicle with two axles each exerting a
vertical load of 20 kN on the truss. The distance between the two axles is 3 meters.

Solution:
Calculate the force in members 𝐶𝐶𝐶𝐶,𝐺𝐺𝐺𝐺, and 𝐷𝐷𝐷𝐷 due to a unit load at joints 𝐵𝐵,𝐶𝐶, and 𝐷𝐷.
Place a unit load at 𝐵𝐵 and determine the support reactions:

Cut the truss through members 𝐶𝐶𝐶𝐶,𝐶𝐶𝐶𝐶, and 𝐺𝐺𝐺𝐺, then write and solve the static equilibirum
equations for the force in memebrs 𝐶𝐶𝐶𝐶 and 𝐺𝐺𝐺𝐺.

� 𝑀𝑀𝐶𝐶 = 0
1 2
−3𝐹𝐹𝐺𝐺𝐺𝐺 + 8 � � = 0 ⇒ 𝐹𝐹𝐺𝐺𝐺𝐺 =
4 3

� 𝑀𝑀𝐻𝐻 = 0
1 1
3𝐹𝐹𝐶𝐶𝐶𝐶 + 4 � � = 0 ⇒ 𝐹𝐹𝐶𝐶𝐶𝐶 = −
4 3

Dr. Structure P a g e |1
Structural Analysis I Truss Influence Line (Moving Load) Exercise Problem Solutions

Use joint D to determine the force in member 𝐷𝐷𝐷𝐷.

� 𝐹𝐹𝑌𝑌 = 0 ⇒ 𝐹𝐹𝐷𝐷𝐷𝐷 = 0

Place a unit load at 𝐶𝐶 and determine the support reactions:

Cut the truss through members 𝐶𝐶𝐶𝐶,𝐶𝐶𝐶𝐶 and 𝐺𝐺𝐺𝐺, then write and solve the static equilibirum
equations for the force in memebrs 𝐶𝐶𝐶𝐶 and 𝐺𝐺𝐺𝐺.

� 𝑀𝑀𝐶𝐶 = 0
1 4
−3𝐹𝐹𝐺𝐺𝐺𝐺 + 8 � � = 0 ⇒ 𝐹𝐹𝐺𝐺𝐺𝐺 =
2 3

� 𝑀𝑀𝐻𝐻 = 0
1 2
3𝐹𝐹𝐶𝐶𝐶𝐶 + 4 � � = 0 ⇒ 𝐹𝐹𝐶𝐶𝐶𝐶 = −
2 3

Use joint D to determine the force in member 𝐷𝐷𝐷𝐷.

� 𝐹𝐹𝑌𝑌 = 0 ⇒ 𝐹𝐹𝐷𝐷𝐷𝐷 = 0

Dr. Structure P a g e |2
Structural Analysis I Truss Influence Line (Moving Load) Exercise Problem Solutions

Place a unit load at 𝐷𝐷 and determine the support reactions:

Cut the truss through members 𝐶𝐶𝐶𝐶,𝐶𝐶𝐶𝐶 and 𝐺𝐺𝐺𝐺, then write and solve the static equilibirum
equations for the force in memebrs 𝐶𝐶𝐶𝐶 and 𝐺𝐺𝐺𝐺.

� 𝑀𝑀𝐶𝐶 = 0
3 2
−3𝐹𝐹𝐺𝐺𝐺𝐺 + 8 � � − 4(1) = 0 ⇒ 𝐹𝐹𝐺𝐺𝐺𝐺 =
4 3

� 𝑀𝑀𝐻𝐻 = 0
3
3𝐹𝐹𝐶𝐶𝐶𝐶 + 4 � � = 0 ⇒ 𝐹𝐹𝐶𝐶𝐶𝐶 = −1
4

Use joint D to determine the force in member 𝐷𝐷𝐷𝐷.

� 𝐹𝐹𝑌𝑌 = 0 ⇒ 𝐹𝐹𝐷𝐷𝐷𝐷 + 1 = 0 ⇒ 𝐹𝐹𝐷𝐷𝐷𝐷 = −1

Dr. Structure P a g e |3
Structural Analysis I Truss Influence Line (Moving Load) Exercise Problem Solutions

Draw influence line for member 𝐺𝐺𝐺𝐺:

Note: The member is always in tension, no compressive force develops in the member.
Draw influence line for member 𝐶𝐶𝐶𝐶:

Note: The member is always in compression, no tensile force develops in the member.

Dr. Structure P a g e |4
Structural Analysis I Truss Influence Line (Moving Load) Exercise Problem Solutions

Draw influence line for member 𝐷𝐷𝐷𝐷:

Note: The member is always in compression, no tensile force develops in the member.
Determine the maximum compressive force in 𝐶𝐶𝐶𝐶:

1 ℎ 3
= ⇒ℎ=
12 9 4
3
𝐹𝐹𝐶𝐶𝐶𝐶 = −20(1) − 20 � � ⇒ 𝐹𝐹𝐶𝐶𝐶𝐶 = −35 kN max. compression
4
Maximum tensile force in CD: F𝐶𝐶𝐶𝐶 = 0

Dr. Structure P a g e |5
Structural Analysis I Truss Influence Line (Moving Load) Exercise Problem Solutions

Determine the maximum tensile force in 𝐺𝐺𝐺𝐺:

ℎ 4/3 5
= ⇒ℎ=
5 8 6
4 5 130
𝐹𝐹𝐺𝐺𝐺𝐺 = 20 � � + 20 � � ⇒ 𝐹𝐹𝐺𝐺𝐺𝐺 = kN max. tension
3 6 3
Maximum compressive force in GH: F𝐺𝐺𝐺𝐺 = 0

Determine the maximum compressive force in 𝐷𝐷𝐷𝐷:

Dr. Structure P a g e |6
Structural Analysis I Truss Influence Line (Moving Load) Exercise Problem Solutions

1 ℎ 1
= ⇒ℎ=
4 1 4
1
𝐹𝐹𝐷𝐷𝐷𝐷 = −20(1) − 20 � � ⇒ 𝐹𝐹𝐷𝐷𝐷𝐷 = −25 kN max. compression
4
Maximum tensile force in DH: F𝐷𝐷𝐷𝐷 = 0

Dr. Structure P a g e |7
Structural Analysis I Truss Influence Line (Moving Load) Exercise Problem Solutions

Problem 2: Determine the maximum tensile and compressive force in member 𝐴𝐴𝐴𝐴 due to the
weight of the moving vehicle shown below.

Solution:
Draw the influence line for member 𝐴𝐴𝐴𝐴:
When the unit load is at either the pin or the roller support, the force in member AB is zero.

Dr. Structure P a g e |8
Structural Analysis I Truss Influence Line (Moving Load) Exercise Problem Solutions

Place the unit load at 𝐵𝐵 and determine the force in member AB:

3
� 𝐹𝐹𝑌𝑌 = 0 ⇒ 2 − 1 − 𝐹𝐹 = 0
5 𝐴𝐴𝐴𝐴

5
𝐹𝐹𝐴𝐴𝐴𝐴 =
3

Place the unit load at 𝐶𝐶 and determine the force in member AB:

Dr. Structure P a g e |9
Structural Analysis I Truss Influence Line (Moving Load) Exercise Problem Solutions

3 3 3
� 𝐹𝐹𝑌𝑌 = 0 ⇒ − − 𝐹𝐹 = 0
2 4 5 𝐴𝐴𝐴𝐴

5
𝐹𝐹𝐴𝐴𝐴𝐴 =
4

Place the unit load at 𝐷𝐷 and determine the force in member AB:

1 3
� 𝐹𝐹𝑌𝑌 = 0 ⇒ 1 − − 𝐹𝐹𝐴𝐴𝐴𝐴 = 0
2 5

5
𝐹𝐹𝐴𝐴𝐴𝐴 =
6

Place the unit load at 𝐸𝐸 and determine the force in member AB:

Dr. Structure P a g e | 10
Structural Analysis I Truss Influence Line (Moving Load) Exercise Problem Solutions

1 1 3
� 𝐹𝐹𝑌𝑌 = 0 ⇒ − − 𝐹𝐹 = 0
2 4 5 𝐴𝐴𝐴𝐴

5
𝐹𝐹𝐴𝐴𝐴𝐴 =
12

Draw the influence line for AB and determine the force in member AB:

Loading Case 1:

5 5 125 105
𝐹𝐹𝐴𝐴𝐴𝐴 = 30 � � + 40 � � + 60 � � + 30 � � ⇒ 𝐹𝐹𝐴𝐴𝐴𝐴 = 215 kN
4 3 96 96

Dr. Structure P a g e | 11
Structural Analysis I Truss Influence Line (Moving Load) Exercise Problem Solutions

Loading Case 2:

5 25 115 95
𝐹𝐹𝐴𝐴𝐴𝐴 = 30 � � + 40 � � + 60 � � + 30 � � ⇒ 𝐹𝐹𝐴𝐴𝐴𝐴 = 214 kN
3 16 96 96

Loading Case 3:

5 35 35 95
𝐹𝐹𝐴𝐴𝐴𝐴 = 30 � � + 60 � � + 40 � � + 30 � � ⇒ 𝐹𝐹𝐴𝐴𝐴𝐴 = 211 kN
3 24 32 96

Dr. Structure P a g e | 12
Structural Analysis I Truss Influence Line (Moving Load) Exercise Problem Solutions

Loading Case 4:

5 5 125 115
𝐹𝐹𝐴𝐴𝐴𝐴 = 30 � � + 60 � � + 40 � � + 30 � � ⇒ 𝐹𝐹𝐴𝐴𝐴𝐴 = 213 kN
6 3 96 96

A comparison of the four loading cases reveals that the maximum tensile force in 𝐴𝐴𝐴𝐴 occurs in
Loading Case 4:

Maximum Tensile Force: 𝐹𝐹𝐴𝐴𝐴𝐴 = 226 kN


Maximum Compressive Force: 𝐹𝐹𝐴𝐴𝐴𝐴 = 0

Dr. Structure P a g e | 13
Lecture 6.6
Truss Influence Lines (Moving Distributed Load)

This lecture focuses on the analysis of trusses subject to moving uniformly distributed loads.

Consider a railroad truss bridge like the one shown in Figure 6.154. The weight of the train that
uses the bridge can be viewed as a variable length uniformly distributed load. The load has a
variable length because the overall length of the train itself is not fixed.

Figure 6.154: A railroad truss bridge subjected to a uniformly distributed load

In the figure above, three truss members have been highlighted: one member along the top
chord, one diagonal member, and one member along the bottom chord. We are going to
determine the maximum compressive and tensile force in each of these members as the train
moves across the bridge.

The bridge is assumed to be simply supported with an overall length of 55 meters, as shown in
Figure 6.155.

Figure 6.155: A simply supported bridge truss

Knowing the dimensions of the truss, we are now ready to perform the required analysis. We
start by making an important observation: Regardless of the position and length of the
distributed load applied to the bridge, the top chord of the truss will always be in compression,

Chapter 6: Influence Lines Page 6.62


while the bottom chord of the truss will always be in tension. The diagonal members, however,
could be either in compression or in tension, depending on the position of the load.

Figure 6.156: The deformed shape of a truss bridge under a train load

In order to determine the maximum axial force in each of the three members, we can draw the
members’ influence lines. This involves placing a unit load at each of the nodes along the bottom
chord of the truss, where the railroad track rests, and calculating the axial force in each member.

When the unit load is placed at the first inner node from the right end of the truss, the left and
right support reactions become 1/10 and 9/10, respectively, as shown in Figure 6.157.

Figure 6.157: The support reactions of a simply supported truss due to a unit load placed at the first
interior node (from right)

Since we are interested in calculating axial force in members AC, BC, and BD, we can cut the
truss as depicted in Figure 6.158, dividing it into two substructures.

Figure 6.158: The free-body diagrams of a truss subjected to a unit load placed at the first interior node
(from right)

Chapter 6: Influence Lines Page 6.63


We can use either of the two free-body diagrams shown in Figure 6.158 to determine the three
unknown member forces (𝐹𝐴𝐶 , 𝐹𝐵𝐶 , and 𝐹𝐵𝐷 ).

Let’s use the free-body diagram of the left side of the truss to calculate the unknown member
forces. In order to write the static equilibrium equations, we need to know the angle that
Member BC makes with the horizontal axis. This angle can be easily determined using basic
trigonometry, as shown in Figure 6.159. The angle is calculated to be approximately 72 degrees.

𝑡𝑎𝑛(𝑎) = 8.5⁄2.75

𝑎 ≈ 72ᵒ

Figure 6.159: The free-body diagram of the left segment of a truss subjected to a unit load placed at the
first interior node (from right)

Since 𝐹𝐵𝐶 is the only unknown force with a y component (see Figure 6.159), we can determine its
magnitude using the following equilibrium equation:

∑ 𝐹𝑦 = (1⁄10) + 𝐹𝐵𝐶 𝑠𝑖𝑛(72) = 0 [6.21]

Solving Equation [6.21] for the unknown force, we get 𝐹𝐵𝐶 = −0.105. To determine 𝐹𝐴𝐶 , we can
write the moment equilibrium equation about Node B:

∑ 𝑀 = 27.5(1⁄10) + 8.5𝐹𝐴𝐶 = 0 [6.22]


𝐵

Solving Equation [6.22] for the unknown force, we get 𝐹𝐴𝐶 = −0.324. Using the moment
equilibrium equation about Node C, we can determine 𝐹𝐵𝐷 :

∑ 𝑀 = 30.25(1⁄10) − 8.5𝐹𝐵𝐷 = 0 [6.23]


𝐶
Solving Equation [6.23] for the unknown force, we get 𝐹𝐵𝐷 = 0.356. Let’s tabulate the computed
member forces in Table 6.4.

Chapter 6: Influence Lines Page 6.64


Unit Load Location (from right) 𝐹𝐵𝐶 𝐹𝐴𝐶 𝐹𝐵𝐷
Roller support 0 0 0
First inner node −0.105 −0.324 0.356

Table 6.4: Axial forces in three members of a truss subjected to a unit load at its first interior node (from
right)

Moving the unit load to the next node, we can calculate the support reactions at the left and
right support to equal 2/10 and 8/10, respectively. Figure 6.160 shows the internal forces in
members AC, BC, and BD, and the free-body diagram of the left segment of the truss.

Figure 6.160: The free-body diagram of the left segment of a truss subjected to a unit load placed at the
second interior node (from right)

Using the free-body diagram shown in Figure 6.160, the following equilibrium equations can be
formulated:

∑ 𝐹𝑦 = (2⁄10) + 𝐹𝐵𝐶 𝑠𝑖𝑛(72) = 0 [6.24]

∑ 𝑀 = 27.5(2⁄10) + 8.5𝐹𝐴𝐶 = 0 [6.25]


𝐵

∑ 𝑀 = 30.25(2⁄10) − 8.5𝐹𝐵𝐷 = 0 [6.26]


𝐶
Solving Equations [6.24], [6.25], and [6.26], we get 𝐹𝐵𝐶 = −0.210, 𝐹𝐴𝐶 = −0.648, and 𝐹𝐵𝐷 =
0.712.

Chapter 6: Influence Lines Page 6.65


The updated table of member axial forces is shown in Table 6.5 in.

Unit Load Location (from right) 𝐹𝐵𝐶 𝐹𝐴𝐶 𝐹𝐵𝐷


Roller support 0 0 0
First inner node −0.105 −0.324 0.356
Second inner node −0.210 −0.648 0.712

Table 6.5: Axial forces in three members of a truss subjected to a unit load at its second interior node
(from right)

We continue moving the unit load along the bottom chord of the truss. In each step, we
determine the support reactions, we cut the truss through members AC, BC, and BD, and then
we write and solve the static equilibrium equations for the unknown member forces.

When the unit load is placed at the third interior node from the right end of the truss, the free-
body diagram in Figure 6.161 for the left segment of the structure can be drawn.

Figure 6.161: The free-body diagram of the left segment of a truss subjected to a unit load placed at the
third interior node (from right)

Using the free-body diagram shown in Figure 6.161, the following equilibrium equations can be
written:

∑ 𝐹𝑦 = (3⁄10) + 𝐹𝐵𝐶 𝑠𝑖𝑛(72) = 0 [6.27]

∑ 𝑀 = 27.5(3⁄10) + 8.5𝐹𝐴𝐶 = 0 [6.28]


𝐵

∑ 𝑀 = 30.25(3⁄10) − 8.5𝐹𝐵𝐷 = 0 [6.29]


𝐶
Solving Equations [6.27] through [6.29] for the unknown member forces, we get 𝐹𝐵𝐶 = −0.315,
𝐹𝐴𝐶 = −0.971, and 𝐹𝐵𝐷 = 1.068. The updated table of member axial forces becomes as shown in
Table 6.6:

Chapter 6: Influence Lines Page 6.66


Unit Load Location (from right) 𝐹𝐵𝐶 𝐹𝐴𝐶 𝐹𝐵𝐷
Roller support 0 0 0
First inner node −0.105 −0.324 0.356
Second inner node −0.210 −0.648 0.712
Third inner node −0.315 −0.971 1.068

Table 6.6: Axial forces in three members of a truss subjected to a unit load at its third interior node
(from right)

When the unit load is placed at the fourth interior node from the right end of the truss, the free-
body diagram in Figure 6.162 for the left segment of the structure can be drawn.

Figure 6.162: The free-body diagram of the left segment of a truss subjected to a unit load placed at the
fourth interior node (from right)

Using the free-body diagram shown in Figure 6.162, the following equilibrium equations can be
written:

∑ 𝐹𝑦 = (4⁄10) + 𝐹𝐵𝐶 𝑠𝑖𝑛(72) = 0 [6.30]

∑ 𝑀 = 27.5(4⁄10) + 8.5𝐹𝐴𝐶 = 0 [6.31]


𝐵

∑ 𝑀 = 30.25(4⁄10) − 8.5𝐹𝐵𝐷 = 0 [6.32]


𝐶

Solving Equations [6.30] through [6.32] for the unknown member forces, we get 𝐹𝐵𝐶 = −0.421,
𝐹𝐴𝐶 = −1.294, and 𝐹𝐵𝐷 = 1.424. The updated table of member axial forces becomes as shown in
Table 6.7:

Unit Load Location (from right) 𝐹𝐵𝐶 𝐹𝐴𝐶 𝐹𝐵𝐷


Roller support 0 0 0
First inner node −0.105 −0.324 0.356
Second inner node −0.210 −0.648 0.712
Third inner node −0.315 −0.971 1.068
Fourth inner node −0.421 −1.294 1.424

Chapter 6: Influence Lines Page 6.67


Table 6.7: Axial forces in three members of a truss subjected to a unit load at its fourth interior node
(from right)

When the unit load is placed at the fifth interior node from the right end of the truss (at Node
B), the free-body diagram in Figure 6.163 for the left segment of the structure can be drawn.

Figure 6.163: The free-body diagram of the left segment of a truss subjected to a unit load placed at the
fifth interior node (from right)

Using the free-body diagram shown in Figure 6.163, the following equilibrium equations can be
written:

∑ 𝐹𝑦 = (5⁄10) − 1 + 𝐹𝐵𝐶 𝑠𝑖𝑛(72) = 0 [6.33]

∑ 𝑀 = 27.5(5⁄10) + 8.5𝐹𝐴𝐶 = 0 [6.34]


𝐵

∑ 𝑀 = 30.25(5⁄10) − 2.75(1) − 8.5𝐹𝐵𝐷 = 0 [6.35]


𝐶

Solving Equations [6.33] through [6.35] for the unknown member forces, we get 𝐹𝐵𝐶 = 0.526,
𝐹𝐴𝐶 = −1.618, and 𝐹𝐵𝐷 = 1.456. The updated table of member axial forces becomes as shown in
Table 6.8:

Unit Load Location (from right) 𝐹𝐵𝐶 𝐹𝐴𝐶 𝐹𝐵𝐷


Roller support 0 0 0
First inner node −0.105 −0.324 0.356
Second inner node −0.210 −0.648 0.712
Third inner node −0.315 −0.971 1.068
Fourth inner node −0.421 −1.294 1.424
Fifth inner node 0.526 −1.618 1.456

Table 6.8: Axial forces in three members of a truss subjected to a unit load at its fifth interior node (from
right)

Note the change in the sign of the axial force in diagonal Member BC. The force has changed
sign from negative (compression) to positive (tension). When the unit load moves to the left of
Node D, the member goes from a compressive state to a tensile state.

Chapter 6: Influence Lines Page 6.68


When the unit load is placed at the sixth interior node from the right end of the truss, the free-
body diagram in Figure 6.164 for the left segment of the structure can be drawn.

Figure 6.164: The free-body diagram of the left segment of a truss subjected to a unit load placed at the
sixth interior node (from right)

Using the free-body diagram shown in Figure 6.164, the following equilibrium equations can be
written:

∑ 𝐹𝑦 = (6⁄10) − 1 + 𝐹𝐵𝐶 𝑠𝑖𝑛(72) = 0 [6.36]

∑ 𝑀 = 27.5(6⁄10) − (27.5 − 22)(1) + 8.5𝐹𝐴𝐶 = 0


[6.37]
𝐵

∑ 𝑀 = 30.25(6⁄10) − (30.25 − 22)(1) − 8.5𝐹𝐵𝐷 = 0


[6.38]
𝐶

Solving Equations [6.36] through [6.38] for the unknown member forces, we get 𝐹𝐵𝐶 = 0.421,
𝐹𝐴𝐶 = −1.294, and 𝐹𝐵𝐷 = 1.165. The updated table of member axial forces becomes as shown in
Table 6.9:

Unit Load Location (from right) 𝐹𝐵𝐶 𝐹𝐴𝐶 𝐹𝐵𝐷


Roller support 0 0 0
First inner node −0.105 −0.324 0.356
Second inner node −0.210 −0.648 0.712
Third inner node −0.315 −0.971 1.068
Fourth inner node −0.421 −1.294 1.424
Fifth inner node 0.526 −1.618 1.456
Sixth inner node 0.421 −1.294 1.165

Table 6.9: Axial forces in three members of a truss subjected to a unit load at its sixth interior node
(from right)

Chapter 6: Influence Lines Page 6.69


When the unit load is placed at the seventh interior node from the right end of the truss, the
free-body diagram in Figure 6.165 for the left segment of the structure can be drawn.

Figure 6.165: The free-body diagram of the left segment of a truss subjected to a unit load placed at the
seventh interior node (from right)

Using the free-body diagram shown in Figure 6.165, the following equilibrium equations can be
written:

∑ 𝐹𝑦 = (7⁄10) − 1 + 𝐹𝐵𝐶 sin(72) = 0 [6.39]

∑ 𝑀 = 27.5(7⁄10) − (27.5 − 16.5)(1) + 8.5𝐹𝐴𝐶 = 0 [6.40]


𝐵

∑ 𝑀 = 30.25(7⁄10) − (30.25 − 16.5)(1) − 8.5𝐹𝐵𝐷 = 0 [6.41]


𝐶

Solving Equations [6.39] through [6.41] for the unknown member forces, we get 𝐹𝐵𝐶 = 0.315,
𝐹𝐴𝐶 = −0.971, and 𝐹𝐵𝐷 = 0.874. The updated table of member axial forces becomes as shown
Table 6.10:

Unit Load Location (from right) 𝐹𝐵𝐶 𝐹𝐴𝐶 𝐹𝐵𝐷


Roller support 0 0 0
First inner node −0.105 −0.324 0.356
Second inner node −0.210 −0.648 0.712
Third inner node −0.315 −0.971 1.068
Fourth inner node −0.421 −1.294 1.424
Fifth inner node 0.526 −1.618 1.456
Sixth inner node 0.421 −1.294 1.165
Seventh inner node 0.315 −0.971 0.874

Table 6.10: Axial forces in three members of a truss subjected to a unit load at its seventh interior node
(from right)

Chapter 6: Influence Lines Page 6.70


When the unit load is placed at the eighth interior node from the right end of the truss, the free-
body diagram in Figure 6.166 for the left segment of the structure can be drawn.

Figure 6.166: The free-body diagram of the left segment of a truss subjected to a unit load placed at the
eighth interior node (from right)

Using the free-body diagram shown in Figure 6.166, the following equilibrium equations can be
written:

∑ 𝐹𝑦 = (8⁄10) − 1 + 𝐹𝐵𝐶 sin(72) = 0 [6.42]

∑ 𝑀 = 27.5(8⁄10) − (27.5 − 11.0)(1) + 8.5𝐹𝐴𝐶 = 0 [6.43]


𝐵

∑ 𝑀 = 30.25(8⁄10) − (30.25 − 11.0)(1) − 8.5𝐹𝐵𝐷 = 0 [6.44]


𝐶

Solving Equations [6.42] through [6.44] for the unknown member forces, we get 𝐹𝐵𝐶 = 0.210,
𝐹𝐴𝐶 = −0.647, and 𝐹𝐵𝐷 = 0.582. The updated table of member axial forces becomes as shown in
Table 6.11:

Unit Load Location (from right) 𝐹𝐵𝐶 𝐹𝐴𝐶 𝐹𝐵𝐷


Roller support 0 0 0
First inner node −0.105 −0.324 0.356
Second inner node −0.210 −0.648 0.712
Third inner node −0.315 −0.971 1.068
Fourth inner node −0.421 −1.294 1.424
Fifth inner node 0.526 −1.618 1.456
Sixth inner node 0.421 −1.294 1.165
Seventh inner node 0.315 −0.971 0.874
Eighth inner node 0.210 −0.647 0.582

Table 6.11: Axial forces in three members of a truss subjected to a unit load at its eighth interior node
(from right)

Chapter 6: Influence Lines Page 6.71


And, when the unit load is placed at the ninth interior node from the right end of the truss, the
free-body diagram in Figure 6.167 for the left segment of the structure can be drawn.

Figure 6.167: The free-body diagram of the left segment of a truss subjected to a unit load placed at the
ninth interior node (from right)

Using the free-body diagram shown in Figure 6.167, the following equilibrium equations can be
written:

∑ 𝐹𝑦 = (9⁄10) − 1 + 𝐹𝐵𝐶 𝑠𝑖𝑛(72) = 0 [6.45]

∑ 𝑀 = 27.5(9⁄10) − (27.5 − 5.5)(1) + 8.5𝐹𝐴𝐶 = 0 [6.46]


𝐵

∑ 𝑀 = 30.25(9⁄10) − (30.25 − 5.5)(1) − 8.5𝐹𝐵𝐷 = 0 [6.47]


𝐶

Solving Equations [6.45] through [6.47] for the unknown member forces, we get 𝐹𝐵𝐶 = 0.105,
𝐹𝐴𝐶 = −0.324, and 𝐹𝐵𝐷 = 0.291. The updated table of member axial forces becomes as shown in
Table 6.12:

Unit Load Location (from right) 𝐹𝐵𝐶 𝐹𝐴𝐶 𝐹𝐵𝐷


Roller support 0 0 0
First inner node −0.105 −0.324 0.356
Second inner node −0.210 −0.648 0.712
Third inner node −0.315 −0.971 1.068
Fourth inner node −0.421 −1.294 1.424
Fifth inner node 0.526 −1.618 1.456
Sixth inner node 0.421 −1.294 1.165
Seventh inner node 0.315 −0.971 0.874
Eighth inner node 0.210 −0.647 0.582
Ninth inner node 0.105 −0.324 0.291
Pin support 0 0 0

Table 6.12: Axial forces in three members of a truss subjected to a unit load at its ninth interior node
(from right)

Chapter 6: Influence Lines Page 6.72


Note that the last row in Table 6.12 shows the member forces when the unit load is at located at
the pin support at the left end of the bridge.

Plotting the axial forces tabulated in Table 6.12, we can create the three influence lines for
Member AC, Member BC, and Member BD (see Figure 6.168).

Figure 6.168: Influence lines for three truss members in a bridge

We are now ready to determine the maximum compressive and tensile force in each truss
member due to the distributed train load.

Suppose the bridge is expected to carry train cars with a length of 12 meters each, and capable of
exerting a maximum load of 40 kN/m on the track. Given that each of the two side trusses
carries half of the total train load, our uniformly distributed load for each truss is 20 kN/m. We
also need to consider that the length of the train load could vary from 12 meters (the length of a
single car) to 55 meters (the overall length of the bridge). Figure 6.169 illustrates the uniformly
distributed load of varying length.

Chapter 6: Influence Lines Page 6.73


Figure 6.169: Distributed loads acting on a bridge side truss

Let’s go through the steps for determining the maximum forces in each of our three target truss
members with the given loading information.

Member BD: The influence line for Member BD is shown in Figure 6.170. Since the member lies
along the bottom chord of the truss, it remains in tension regardless of the position of the load.
This fact is reflected in the shape of the member’s influence line. That is, the entire diagram lies
above the 𝑥-axis.

Figure 6.170: The influence line for Member BD along the bottom chord of a bridge truss

To determine the maximum tensile force in the member, we simply multiply the distributed load
magnitude by the (positive) area under the influence line. Since the influence line consists of

Chapter 6: Influence Lines Page 6.74


straight lines, we can determine the area beneath it using basic geometry. The entire area can be
divided into two triangles (one at each end) and eight trapezoids. We can easily determine the
area of each basic shape, and then add up the individual areas to obtain the entire area under
the influence line. Doing so, we find that it equals 43.59.

So, for a uniformly distributed load of 20 kN/m, the maximum tensile force in the member can
be calculated by multiplying 20 by 43.59, as shown below.

𝑚𝑎𝑥. 𝑡𝑒𝑛𝑠𝑖𝑙𝑒 𝑓𝑜𝑟𝑐𝑒 = (20)(43.59) ≅ 872 𝑘𝑁 [6.48]

This maximum axial force develops in the member when the train covers the entire length of the
bridge. Since the influence line has no negative region, no compressive force develops in the
member, regardless of the length and position of the train.

Member BC: The influence line for the inclined Member BC has a positive area and a negative
area, as shown in Figure 6.171. When the distributed load covers the entire positive area of the
influence line, the tensile force in the member reaches its maximum value. When the load covers
the entire negative area of the influence line, the compressive force in the member reaches its
maximum value.

To determine these maximum values, therefore, we need to calculate the areas associated with
the positive and negative regions of the influence line.

Figure 6.171: The influence line for diagonal Member BC in a bridge truss

As shown in Figure 6.171, a change in sign occurs at the middle segment of the influence line.
This segment of the diagram is more closely examined in Figure 6.172.

Chapter 6: Influence Lines Page 6.75


Figure 6.172: An influence line segment consisting of positive and negative regions

Because Figure 6.172 shows two similar triangles, we know that we can use the base-to-height
ratio equality of the two triangles to determine distance x, as shown below.

𝑥 ⁄4.21 = (5.5 − 𝑥)⁄0.526 [6.49]

Solving Equation [6.49] for x, we get 𝑥 = 2.44 meters. Therefore, the distance from the left end
of the truss to the point where the influence line crosses the x-axis equals 6 × 5.5 − 2.44 = 30.56
meters.

Figure 6.173 shows the positive region of the influence line where the distributed load must be
placed in order to produce the maximum tensile force in the member.

Figure 6.173: The positive region of the influence line for a diagonal truss member subjected to a
distributed load

The area under the positive region of the influence line shown in Figure 6.173 can be calculated
to be 8.04. This is the area of the triangle shown in blue. Therefore, the maximum tensile force
in the member becomes:

𝑚𝑎𝑥. 𝑡𝑒𝑛𝑠𝑖𝑙𝑒 𝑓𝑜𝑟𝑐𝑒 = (20)(8.04) ≅ 161 𝑘𝑁 [6.50]

To determine the maximum compressive force in the member, we place the distributed load
over the negative region of the influence line, as shown in Figure 6.174.

Figure 6.174: The negative region of the influence line for a diagonal truss member subjected to a
distributed load

Chapter 6: Influence Lines Page 6.76


The area under the influence line’s negative region can be calculated to be 5.14. This is the area
of the triangle shown in red. Therefore, the maximum compressive force in the member can be
computed as shown:

𝑚𝑎𝑥. 𝑐𝑜𝑚𝑝𝑟𝑒𝑠𝑠𝑖𝑣𝑒 𝑓𝑜𝑟𝑐𝑒 = (20)(5.14) ≅ 103 𝑘𝑁 [6.51]

Member AC: The influence line for Member AC has no positive region; the entire diagram lies
below the x-axis (see Figure 6.175). Therefore, the member carries compressive force only.

Figure 6.175: The influence line for Member AC along the top chord of a bridge truss

The maximum compressive force in the member can be determined by multiplying the load
magnitude (20 kN/m) by the area under the influence line. Since the diagram forms a triangle
with a peak value of 1.618, its area is 55 × 1.618⁄2 = 44.495. Therefore, the maximum force in
the member can be written as follows:

𝑚𝑎𝑥. 𝑐𝑜𝑚𝑝𝑟𝑒𝑠𝑠𝑖𝑣𝑒 𝑓𝑜𝑟𝑐𝑒 = (20)(44.50) ≅ 890 𝑘𝑁 [6.52]

The maximum compressive and tensile forces in the remaining truss members can be
determined in a similar manner.

We’ll continue our discussion on this topic in the next lecture.

Chapter 6: Influence Lines Page 6.77


Lecture 6.7
Analysis of a Short-Span Highway Bridge

This lecture focuses on the use of influence lines for the analysis of a short-span highway bridge.

Consider the bridge shown in Figure 6.176. Suppose we wish to determine the maximum
reaction force, shear force, and bending moment in the structure due to a combination of
permanent and transient loads.

Figure 6.176: A short-span highway bridge

As can be seen in Figure 6.177, the bridge deck rests on four identical girders. For this analysis,
we can assume that the bridge loads are distributed equally among the girders. Knowing this, we
can determine the extreme internal forces in a typical girder.

Figure 6.177: The girder system for a short-span bridge

Chapter 6: Influence Lines Page 6.78


Highway bridges are generally designed to carry permanent and transient loads. For example,
the weight of the bridge deck is considered a permanent load; it is not going to change
significantly during the life of the structure. On the other hand, vehicles exert transient loads on
the bridge. A transient load changes position and magnitude as a function of time.

In this lecture, we are going to consider two permanent loads: the weight of the bridge deck and
the weight of the girder. For transient loads, we will use a widely accepted load pattern referred
to as HL-93 loading, given by the American Association of State Highway and Transportation
Officials (AASHTO).

Let’s start the analysis by calculating the magnitude of the permanent loads.

The bridge deck is made of reinforced concrete with an average specific weight of 25 kN/m3. As
shown in Figure 6.178, the deck has a width of 14 meters and an average thickness of 0.3 meters.

Figure 6.178: The width and thickness of a short-span highway bridge

The weight per linear meter of the deck is the product of the specific weight of the deck (25
kN/m3), the width of the deck (14 m), and the thickness of the deck (0.3 m), or,

(25 𝑘𝑁/𝑚3 )(14 𝑚)(0.3 𝑚) = 105 𝑘𝑁/𝑚 [6.53]

As mentioned, we can assume that this weight is distributed equally among the four girders.
Hence, each girder carries a uniformly distributed load of 26.25 kN/m, as depicted in
Figure 6.179.

Chapter 6: Influence Lines Page 6.79


Figure 6.179: The uniformly distributed load due to the weight of concrete on a bridge girder

For this analysis, we can use 90,000 mm2 (0.09 m2) for the cross-sectional area of each steel
girder. Assuming a specific weight of 78 kN/m3 for steel, the self-weight of each girder, viewed as
a uniformly distributed load, becomes:

(78 𝑘𝑁/𝑚3 )(0.09 𝑚2 ) = 7.02 𝑘𝑁/𝑚 [6.54]

Therefore, the total permanent load uniformly distributed over a typical girder can be obtained
by adding the loads due to the calculated weights of concrete and steel.

26.25 + 7.02 = 33.27 𝑘𝑁/𝑚 [6.55]

For transient loading using AASHTO HL-93 specifications, we consider three loads: one
uniformly distributed load and two concentrated loads.

The uniformly distributed load represents a series of hypothetical vehicles moving in tandem.
The magnitude of this load, per AASHTO HL-93, is 9.3 kN/m (see
Figure 6.180). We call this the Design Lane Load.

Figure 6.180: Design Lane Load for a short-span highway bridge

Chapter 6: Influence Lines Page 6.80


In addition, HL-93 requires us to consider two sets of concentrated loads: the Design Truck
Load and the Design Tandem Load. The Design Truck Load consists of three concentrated loads
that represent one load for each axle of a hypothetical three-axle truck. These loads are shown in
Figure 6.181.

Figure 6.181: Design Truck Load for a short-span highway bridge

The Design Tandem Load consists of two concentrated loads, each with a magnitude of 110 kN,
as shown in Figure 6.182.

Figure 6.182: Design Tandem Load for a short-span highway bridge

In applying the transient loads to the bridge, we need to determine the load combination that
produces the maximum internal force in the structure. Herein, we’ll focus on determining the
maximum support reaction, the maximum shear force, and the maximum positive and negative
moments in a typical girder.

Our target girder is shown in Figure 6.183. The girder rests on a pin support at the left end of the
bridge, a roller support at the interior pier, and a roller support at the right end of the bridge.
Furthermore, a splice is at the midpoint of the girder, which we will treat as an internal hinge.

Chapter 6: Influence Lines Page 6.81


Figure 6.183: The supports for a typical short-span bridge girder

To determine the maximum support reaction in the girder due to the permanent and transient
loads, we start by drawing the reaction influence line for each support.

Figure 6.184: The line diagram of a typical short-span bridge girder

Given the line diagram of the girder shown in Figure 6.184, we can draw the influence line for
the reaction force at the pin support at Point A as follows in Figure 6.185:

Figure 6.185: The reaction influence line for the left support of a short-span bridge girder

For the reaction force at the roller at Point B, the influence line can be drawn as shown in Figure
6.186:

Figure 6.186: The reaction influence line for the middle support of a short-span bridge girder

And Figure 6.187 shows the reaction influence line for the roller support at Point C at the right
end of the girder:

Chapter 6: Influence Lines Page 6.82


Figure 6.187: The reaction influence line for the right support of a short-span bridge girder

The reaction influence lines shown in Figure 6.185 through Figure 6.187 can be used to calculate
the support reactions due to the permanent and transient loads.

Let’s start with the influence line for the left support reaction, shown again in Figure 6.188.
Using basic geometry, we can determine height ℎ with the equation 1⁄9 = ℎ⁄3. Solving for ℎ, we
get ℎ = 1⁄3.

Figure 6.188: The reaction influence line for the left support of a short-span bridge girder

Previously, we determined the magnitude of the permanent load uniformly distributed over a
typical girder as 33.27 kN/m. Figure 6.189 shows the uniformly distributed permanent load
along our line diagram.

Figure 6.189: The permanent uniformly distributed load on a typical girder in a short-span bridge

If we multiply this load magnitude by the area under the influence line, we get the reaction force
at the support at Point A due to the permanent load. Note that the influence line consists of a
positive region and a negative region (see Figure 6.190).

Figure 6.190 shows that the area of the positive region in the diagram is (1) (9)⁄2 = 4.5.
Similarly, the area of the negative region in the diagram is 2.5.

Figure 6.190: The areas under the reaction influence line for the left support of a short-span bridge
girder

Chapter 6: Influence Lines Page 6.83


Therefore, the net area under the influence line equals 4.5 − 2.5 = 2. If we multiply the
permanent load magnitude by the total area under the influence line, we get (33.27)(2) =
66.54 𝑘𝑁. This is the reaction force at the left support due to the permanent load.

To determine the reaction force at the interior roller support, we can multiply the magnitude of
the distributed permanent load by the area under the influence line for the support. Figure 6.191
shows the distributed permanent load and the influence line for the middle roller support.

Figure 6.191: The area under the reaction influence line for the interior support of a short-span bridge
girder

Height ℎ in Figure 6.191 can be calculated by equating the height-to-base ratio of the smaller
right triangle to that of the larger right triangle. Doing so, we get 1⁄9 = ℎ⁄12, or ℎ = 4⁄3.
Therefore, the area of the triangle defined by the influence line equals (4⁄3) (24)⁄2 = 16.

The reaction force at the roller at Point B, therefore, can be computed by multiplying 33.27 by
16. So, the reaction force due to the permanent load equals (33.27)(16) = 532.32 𝑘𝑁.

To determine the reaction force at the right support, we start again with the reaction influence
line for the right support. Figure 6.192 shows the distributed permanent load and the influence
line for the right roller support.

Figure 6.192: The area under the reaction influence line for the right support of a short-span bridge
girder

The area under the influence line here equals (1) (12)⁄2 = 6. Therefore, the reaction force at the
roller at Point C due to the permanent load equals (33.27)(6) = 199.62 𝑘𝑁.

Chapter 6: Influence Lines Page 6.84


By comparing the three computed reaction force values (see Figure 6.193), we can conclude that
the maximum reaction force in the beam due to the permanent load occurs at the roller at Point
B.

Figure 6.193: The support reaction forces for a short-span bridge girder

However, our analysis is not complete yet. We also need to include the effect of the transient
loads on the support reactions.

Per AASHTO HL-93, three transient loads need to be considered: the Design Lane Load, the
Design Truck Load, and the Design Tandem Load. These loads are shown in Figure 6.194.

Figure 6.194: Three transient loads for a short-span highway bridge

For design purposes, AASHTO HL-93 requires us to consider two load combinations: the Design
Lane Load plus the Design Truck Load, and the Design Lane Load plus the Design Tandem
Load. These two load combinations are shown in Figure 6.195.

Figure 6.195: Two transient load combinations for the analysis of a short-span bridge girder

The load combination that produces the maximum reaction force governs the design.

Chapter 6: Influence Lines Page 6.85


Before we calculate the reaction forces due to the transient loads, let’s look at how these loads
will impact a single girder in our highway bridge. Consider the following load combination in
Figure 6.196:

Figure 6.196: The Design Lane Load and Design Truck Load combination for the analysis of a
short-span bridge

Since the bridge deck rests on four girders and each traffic lane is supported by two girders, we
can assume that each girder carries half of the transient load, as depicted in Figure 6.197.

Figure 6.197: A transient load distributed equally between two girders in each traffic lane in a short-
span highway bridge

Therefore, we must divide the load magnitudes in half in order to arrive at the design loads that
act on a single girder, as shown in Figure 6.198.

Figure 6.198: The Design Lane Load and Design Truck Load combination for the analysis of a bridge
girder

Chapter 6: Influence Lines Page 6.86


Let’s place these three transient loads on the beam one at a time and calculate the maximum
support reaction that results. We’ll start with the Design Lane Load.

The maximum reaction force at Point A can be attained by placing the distributed Design Lane
Load over the part of the beam that corresponds to the positive region under the influence line,
as shown in Figure 6.199.

Figure 6.199: The reaction influence line for the support at the left end of a girder and the Design Lane
Load pattern that maximizes the reaction force

Similarly, for the reaction force at the roller support at Point B, we stretch the distributed Design
Lane Load over the part of the beam that corresponds to the positive region under the influence
line for the reaction force (see Figure 6.200). This loading pattern yields the maximum reaction
force at the support.

Figure 6.200: The reaction influence line for the interior support in a girder and the Design Lane Load
pattern that maximizes the reaction force

And again, to determine the maximum reaction force at the support at Point C, we place the
distributed Design Lane Load over the positive region of the influence line for the reaction force,
as shown in Figure 6.201.

Figure 6.201: The reaction influence line for the support at the right end of a girder and the Design Lane
Load pattern that maximizes the reaction force

Chapter 6: Influence Lines Page 6.87


A qualitative comparison of the three influence line diagrams and load patterns shown in Figure
6.199 through Figure 6.201 reveals that the interior roller support at Point B carries the largest
force. We can tell this because the influence line for the reaction force at Point B has the largest
positive region.

We can perform the same type of qualitative analysis for the concentrated transient loads. For
the Design Truck Load, the two front axles of the truck are separated by 4.3 meters. The rear
axles can be set apart by 4.3 to 9 meters, as shown in Figure 6.202.

Figure 6.202: The concentrated load series for the AASHTO HL-93 Design Truck Load

Our task is to determine the truck location and orientation that produces the maximum reaction
force in the girder under the Design Truck Load. Figure 6.203 shows the truck locations that
produce a large reaction force at each support.

(a)

(b)

(c)

Chapter 6: Influence Lines Page 6.88


Figure 6.203: The Design Truck Load locations for producing large reaction forces at the supports of a
short-span bridge girder

For the cases shown in Figure 6.203, we need to multiply each concentrated load by the height
of the influence line under the load. We can then add up the three resulting values to get the
magnitude of the reaction force. Given the shapes of the three influence lines, it should be
obvious that the support reaction at Point B is going to be larger than the reaction force at either
Point A or Point C. Therefore, the case given in Figure 6.203(b) governs the design.

In other words, to determine the maximum reaction force in the beam under the Design Truck
Load, we can safely ignore the reactions at Point A and Point C and focus solely on the reaction
force at Point B.

A more detailed representation of Figure 6.203(b) is given in Figure 6.204.

Figure 6.204: The Design Truck Load location for producing the largest reaction force in a short-span
bridge girder

The diagram shown in Figure 6.204 can be used to calculate the maximum reaction force at
Point B. As stated before, this is done by multiplying the magnitude of each concentrated load by
the height of the influence line under the load. The height of the influence line under the front
and rear loads can be determined using basic geometry.

Height ℎ can be calculated as follows:

ℎ⁄7.7 = (4⁄3)⁄12 ⇒ ℎ = 0.86 [6.56]

Similarly, we can calculate height 𝑔 as shown below:

𝑔⁄7.7 = (4⁄3)⁄12 ⇒ 𝑔 = 0.86 [6.57]

Chapter 6: Influence Lines Page 6.89


Then, we can compute the reaction force at Point B due to the Design Truck Load in the
following manner:

(72.5)(0.86) + (72.5)(4⁄3) + (17.5)(0.86) = 173.67 𝑘𝑁 [6.58]

AASHTO HL-93 indicates that multiple trucks can be present on a beam, as long as the tail-to-
nose distance between two consecutive trucks is not less than 15 meters. However, our short-
span bridge cannot accommodate more than one full truck at a time. Therefore, the maximum
reaction force due to the Design Truck Load equals 173.67 kN.

Let’s summarize what we have done so far. We have:

1. Determined that the maximum reaction force in the girder occurs at Point B.
2. Calculated the reaction force at Point B due to the permanent load (532.32 kN).
3. Calculated the reaction force at Point B due to the Design Lane Load (74.4 kN).
4. Calculated the reaction force at Point B due to the Design Truck Load (173.67 kN).
The combined Design Lane Load and Design Truck Load gives a reaction force of 248.07 kN at
the roller support at Point B. But we also need to calculate the reaction force at Point B due to
the Design Tandem Load and decide which load combination governs the design: the Design
Lane Load plus the Design Truck Load, or the Design Lane Load plus the Design Tandem Load.

Per AASHTO HL-93 specifications, the Design Tandem Load consists of two concentrated loads
set 1.2 meters apart. The magnitude of each load per girder equals 55 kN.

To determine the maximum effect of the Design Tandem Load on the reaction force at Point B,
we can place the truck on the bridge as shown in Figure 6.205.

Figure 6.205: The Design Tandem Load location for producing the largest reaction force in a short-span
bridge girder

The resulting reaction force at Point B due to the Design Tandem Load can be calculated as
shown:

55(4⁄3) + 55(1.2) = 139.33 𝑘𝑁 [6.59]

Although AASHTO HL-93 does not specify a minimum distance between trucks for the Design
Tandem Load, we can assume a practical distance for separating two consecutive trucks. In this

Chapter 6: Influence Lines Page 6.90


example we’ll use 10 meters. This means that we can place three trucks on the bridge, as shown
in Figure 6.206.

Figure 6.206: A series of truck Design Tandem Loads for producing the maximum reaction force in a
short-span bridge girder

Using the influence line shown in


Figure 6.206, we can determine the maximum reaction force at the interior support due to a
series of Design Tandem Loads as follows:

55(0.09 + 0.22 + 4⁄3 + 1.2 + 0.09) = 161.33 𝑘𝑁 [6.60]

Since this value is lower than the reaction force caused by the Design Truck Load, the governing
transient load for calculating the maximum reaction force in the bridge is the combination of the
Design Lane Load and the Design Truck Load. This combination equals 173.67 + 74.4 ≅ 248 𝑘𝑁.

Finally, we need to combine the permanent and transient loads using the Load and Resistance
Factor Design (LRFD) load combination equations. For HL-93 loading, we can use the following
equation:

𝐿𝑅𝐹𝐷 𝑙𝑜𝑎𝑑 𝑐𝑜𝑚𝑏𝑖𝑛𝑎𝑡𝑖𝑜𝑛 𝑒𝑞𝑢𝑎𝑡𝑖𝑜𝑛: 1.25 𝑝𝑒𝑟𝑚𝑎𝑛𝑒𝑛𝑡 𝑙𝑜𝑎𝑑 + 1.75 𝑡𝑟𝑎𝑛𝑠𝑖𝑒𝑛𝑡 𝑙𝑜𝑎𝑑 [6.61]

Chapter 6: Influence Lines Page 6.91


Substituting the calculated load magnitudes into Equation [6.61], we get:

1.25(532.32) + 1.75(248.07) ≅ 1100 𝑘𝑁 [6.62]

Therefore, the absolute maximum design reaction force in the bridge, which takes place at the
roller support at Point B, equals 1100 kN.

Next, let’s turn our attention to the shear force analysis of the girder. Suppose we wish to
determine the maximum shear force that could develop at the internal hinge at Point D (see
Figure 6.207).

Figure 6.207: A short-span bridge girder with an internal hinge

We can start by drawing the shear influence line for the hinge, as shown in Figure 6.208.

Figure 6.208: The shear influence line for an internal hinge in a bridge girder

This influence line can be used to calculate the maximum shear force at Point D due to the
permanent and transient loads.

We already know that the permanent uniformly distributed load has a magnitude of 33.27 kN/m.
It is shown in
Figure 6.209. And as can be calculated from the shear influence line diagram also in
Figure 6.209, the area under the influence line equals 6. Therefore, the shear force at Point D
due to the permanent load can be calculated as follows:

(33.27)(6) = 199.62 𝑘𝑁 [6.63]

Figure 6.209: The permanent load and the shear influence line for an internal hinge in a girder

Chapter 6: Influence Lines Page 6.92


For the Design Lane Load, we place the uniformly distributed load of 4.65 kN/m on the part of
the girder associated with the positive area under the influence line (see Figure 6.210).

Figure 6.210: The shear influence line for an internal hinge in a bridge girder and the Design Lane Load
pattern that maximizes the shear force

Using Figure 6.210, we can calculate the shear force at the hinge due to the Design Lane Load as
follows:

(4.65)(6) = 27.90 𝑘𝑁 [6.64]

And the Design Truck Load, when placed on the bridge for maximum effect, yields the following
load pattern shown in Figure 6.211:

Figure 6.211: The Design Truck Load location for producing the largest shear force at an internal hinge
in a bridge girder

Knowing the position of each concentrated load, we can easily calculate the height of the
influence line under the middle and front loads. These heights are shown in Figure 6.211. Hence,
the shear force at the hinge due to the Design Truck Load can be calculated as shown:

(72.5)(1) + (72.5)(0.64) + (17.5)(0.28) = 123.97 𝑘𝑁 [6.65]

Finally, to determine the contribution of the Design Tandem Load to the maximum shear force
at the internal hinge, we can place the load series on the girder for maximum effect. This is
shown in Figure 6.212.

Chapter 6: Influence Lines Page 6.93


Figure 6.212: The Design Tandem Load location for producing the largest shear force at an internal
hinge in a bridge girder

The shear force at the hinge due to the Design Tandem Load can be calculated as follows:

(55)(1) + (55)(0.9) = 104.5 𝑘𝑁 [6.66]

The calculated shear forces at the internal hinge are tabulated in Table 6.13.

Load Type Shear Force Magnitude


Permanent Load 199.62 kN
Design Lane Load 27.90 kN
Design Truck Load 123.97 kN
Design Tandem Load 104.50 kN

Table 6.13: The shear force magnitudes at an internal hinge in a bridge girder due to permanent and
transient loads

Comparing the shear forces due to the Design Tandem Load and the Design Truck Load in Table
6.13, we can conclude that the Design Truck Load governs the design. Therefore, the total shear
force due to transient load is 27.90 + 123.97 = 151.87 𝑘𝑁.

Then, using the LRFD load combination equation (see Equation [6.61]), we can calculate the
maximum design shear force at the hinge as shown below:

1.25(199.62) + 1.75(151.87) = 515.30 𝑘𝑁 [6.67]

Lastly, suppose we want to determine the maximum positive and negative bending moments in
the girder. To start, we can use qualitative reasoning to determine the location of the maximum
positive moment in the girder. In this case, bending moment reaches its maximum positive
value at the midpoint of either Segment AB or Segment DC. These points have been labeled
respectively as points E and F in Figure 6.213.

Figure 6.213: The points of maximum positive bending moment in a bridge girder

We can use the moment influence lines for Point E and Point F to determine which point
controls the design. The moment influence lines for the two points are shown in Figure 6.214.

Chapter 6: Influence Lines Page 6.94


Figure 6.214: The moment influence lines for two midpoints in a bridge girder

The peak value at Point E in Figure 6.214 represents the magnitude of the bending moment at
Point E when the beam is subjected to a unit load placed at Point E, as shown in Figure 6.215.

Figure 6.215: A girder subjected to a unit load and the resulting free-body diagram of its left segment

Using the free-body diagram shown in Figure 6.215, we can calculate the bending moment at
Point E as follows:

𝑀𝐸 = 0.5(4.5) = 2.25 [6.68]

Therefore, the maximum height of the moment influence line for Point E is 2.25, as shown in
Figure 6.216.

Figure 6.216: The moment influence line for the midpoint of a segment of a bridge girder

Height ℎ in Figure 6.216 can be calculated as shown:

ℎ⁄3 = 2.25⁄4.5 ⇒ ℎ = 1.5 [6.69]

We can determine the height of the moment influence line for Point F in a similar manner (see
Figure 6.214). To do so, we place a unit load at Point F, calculate the right support reaction, and
then draw the free-body diagram of the right segment of the girder. The result is shown in
Figure 6.217.

Chapter 6: Influence Lines Page 6.95


Figure 6.217: A girder subjected to a unit load and the resulting free-body diagram of its right segment

Using the free-body diagram shown in Figure 6.217, we can determine the bending moment at
Point F as follows:

𝑀𝐹 = 0.5(6) = 3 [6.70]

Knowing the heights of the two influence lines previously shown in Figure 6.214, we can redraw
them as shown in Figure 6.218:

Figure 6.218: The complete moment influence lines for two midpoints in a bridge girder

The area of the positive region under the moment influence line for Point E is 2.25(9/2) =
10.125. The area of the negative region of this influence line equals 1.5(15/2) = 11.25.
Therefore, the net area becomes 10.125 − 11.25 = −1.125.

For the moment influence line for Point F, the total positive area equals 3(12⁄2) = 18.

Since the influence line with the largest positive area controls the design, we can conclude that
the bending moment in the beam reaches its maximum value at Point F.

Now that we have identified the location of the maximum positive bending moment, we can now
begin calculating the magnitude of the moment due to the permanent and transient loads. To
determine the maximum bending moment in the beam due to the permanent load, we can
multiply the load magnitude by the area under the influence line. Figure 6.219 shows the
uniformly distributed permanent load and the influence line for the moment at Point F.

Chapter 6: Influence Lines Page 6.96


Figure 6.219: The permanent load placement for producing the maximum positive bending moment in
the girder

The maximum moment in the girder due to the permanent load is calculated as follows:

(33.27)(18) = 598.86 𝑘𝑁. 𝑚 [6.71]

Then, to calculate the moment value at Point F due to the Design Lane Load, we load the girder
as shown in Figure 6.220.

Figure 6.220: The Design Lane Load placement for producing the maximum positive bending moment
in a bridge girder

The maximum moment in the girder due to the Design Lane Load becomes:

(4.65)(18) = 83.7 𝑘𝑁. 𝑚 [6.72]

Next, to maximize the effect of the Design Truck Load on the positive bending moment in the
girder, we can place the truck load series on the girder as shown in Figure 6.221:

Chapter 6: Influence Lines Page 6.97


Figure 6.221: The Design Truck Load placement for producing the maximum positive bending moment
in a bridge girder

Using the diagram shown in Figure 6.221, we can calculate the maximum bending moment at
Point F due to the Design Truck Load as follows:

(72.5)(0.85) + (72.5)(3) + (17.5)(0.85) = 294 𝑘𝑁. 𝑚 [6.73]

And finally, for the Design Tandem Load, we can use the following diagrams in Figure 6.222:

Figure 6.222: The Design Tandem Load placement for producing the maximum positive bending
moment in a bridge girder

Using Figure 6.222, we can calculate the maximum bending moment at Point F due to the
Design Tandem Load as follows:

(55)(3) + (55)(2.4) = 297 𝑘𝑁. 𝑚 [6.74]

The calculated positive bending moments in the girder under different load cases are tabulated
in Table 6.14.

Load Type Positive Bending Moment


Magnitude
Permanent Load 598.96 kN·m
Design Lane Load 83.7 kN·m
Design Truck Load 294 kN·m
Design Tandem Load 297 kN·m

Chapter 6: Influence Lines Page 6.98


Table 6.14: The positive bending moment magnitudes in a bridge girder due to permanent and transient
loads

Per Table 6.14, the moment due to the Design Tandem Load is larger than the moment due to
the Design Truck Load. So, the maximum positive bending moment due to the transient loads
becomes 83.7 + 297 = 380.7 𝑘𝑁. 𝑚.

Finally, using the LRFD load combination equation (Equation [6.61]), we can determine the
maximum positive bending moment in the beam as shown:

1.25(598.86) + 1.75(380.7) = 1414.8 kN. 𝑚 [6.75]

We can now repeat essentially the same process that we used to find the maximum positive
bending moment to find the maximum negative bending moment. To determine the maximum
negative bending moment in the girder, we first need to determine its position on the girder.
The two candidates for points at which the maximum negative moment can develop are Point E
and Point B.

The moment influence line for Point E and the areas associated with its positive and negative
regions are shown in Figure 6.223.

Figure 6.223: A moment influence line and the areas associated with its positive and negative regions

The moment influence line for Point B is shown in Figure 6.224.

Figure 6.224: The moment influence line for an interior support in a bridge girder

To determine the height of the influence line shown in Figure 6.224, we can place a unit load at
Point D, and then calculate the bending moment that results at Point B (see Figure 6.225).

Chapter 6: Influence Lines Page 6.99


Figure 6.225: A bridge girder subjected to a unit load at an internal hinge

To calculate the bending moment at Point B, we can isolate Segment AD and draw its free-body
diagram. The free-body diagram is shown in Figure 6.226.

Figure 6.226: A segment of a girder under a unit load at its right end

Using Figure 6.226, we can calculate the bending moment value at Point B as follows:

𝑀𝐵 = (1)(3) = 3 [6.76]

Therefore, the moment influence line for Point B can be shown as follows in Figure 6.227:

Figure 6.227: The complete moment influence line for an interior support in a bridge girder

The area under the influence line shown in Figure 6.227 is 3(15⁄2) = 22.5. Since we can
qualitatively see that this area is larger than the area of the negative region associated with the
moment influence line for Point E (see Figure 6.223), we can conclude that the largest negative
moment in the beam occurs at Point B.

Chapter 6: Influence Lines Page 6.100


To determine the magnitude of the maximum negative moment, we need to load the girder with
the permanent and transient loads. Figure 6.228 shows the permanent load pattern for
producing the maximum negative moment at Point B.

Figure 6.228: The permanent load pattern for producing the maximum negative moment at the interior
support of a bridge girder

From the diagrams shown in Figure 6.228, we can calculate the negative bending moment at
Point B due to the permanent load as follows:

(33.27)(22.5) = 748.58 𝑘𝑁. 𝑚 [6.77]

To determine the negative moment at Point B due to the Design Lane Load, we can load the
beam as shown in Figure 6.229:

Figure 6.229: The Design Lane Load pattern for producing the maximum negative moment at the
interior support of a bridge girder

From the diagrams shown in Figure 6.229, we can calculate the negative bending moment due
to the Design Lane Load at Point B as shown:

(4.65)(22.5) = 104.63 𝑘𝑁. 𝑚 [6.78]

And to maximize the effect of the Design Truck Load on the negative bending moment in the
girder, we can place the truck load series on the girder as shown in Figure 6.230:

Chapter 6: Influence Lines Page 6.101


Figure 6.230: The Design Truck Load placement for producing the maximum negative bending moment
in a bridge girder

The negative bending moment at Point B due to the Design Truck Load shown in Figure 6.230
can be calculated as shown:

(72.5)(3) + (72.5)(1.93) + (17.5)(0.85) = 372.30 𝑘𝑁. 𝑚 [6.79]

And, for the Design Tandem Load, we can use the following diagrams shown in Figure 6.231:

Figure 6.231: The Design Tandem Load placement for producing the maximum negative bending
moment in a bridge girder

The negative bending moment at Point B due to the Design Tandem Load shown in Figure 6.231
can be calculated as shown:

(55)(3) + (55)(2.7) = 313.5 𝑘𝑁. 𝑚 [6.80]

However, using our practical distance of 10 meters between two consecutive trucks, we can see
that for the Design Tandem Load, a second truck can be placed on the girder as shown in Figure
6.232.

Chapter 6: Influence Lines Page 6.102


Figure 6.232: Two Design Tandem Load truck placements for producing the maximum negative
bending moment in a bridge girder

The resulting bending moment at Point B due to the Design Tandem Load series shown in
Figure 6.232 can be calculated as follows:

(55)(3) + (55)(2.7) + 55(0.2) = 324.5 𝑘𝑁. 𝑚 [6.81]

Table 6.15 shows the results of the calculation for determining the maximum negative moment
in the girder.

Negative Bending Moment


Load Type
Magnitude
Permanent Load 748.58 kN·m
Design Lane Load 104.63 kN·m
Design Truck Load 372.30 kN·m
Design Tandem Load 324.50 kN·m

Table 6.15: The negative bending moment magnitudes in a bridge girder due to permanent and
transient loads

Since the bending moment due to the Design Truck Load is larger than the bending moment due
to the Design Tandem Load, we calculate the total negative bending moment due to the
transient load by combining the moment values due to the Design Lane Load and the Design
Truck Load. The total negative bending moment is calculated as shown:

104.63 + 372.30 ≅ 477 𝑘𝑁. 𝑚 [6.82]

Then, using the LRFD load combination equation (Equation [6.61]), we can determine the
maximum negative design bending moment in the girder:

1.25(748.58) + 1.75(476.93) ≅ 1770 kN. m [6.83]

Chapter 6: Influence Lines Page 6.103


In conclusion, for our short-span bridge, the maximum reaction force in each girder develops at
the interior support. The force magnitude is 1100 kN.

The maximum shear force at the hinge in a typical girder is 515 kN.

The maximum positive moment takes place 6 meters away from the right end of the bridge. The
magnitude of the bending moment in each girder is 1415 kN·m.

And, the maximum negative moment occurs at the interior support of each girder. The
magnitude of the moment is 1770 kN·m.

Chapter 6: Influence Lines Page 6.104


Lecture 6.8
Analysis of a Floor Girder

In previous lectures, we presented and illustrated the process of constructing influence lines for
beam and truss bridges subjected to vehicular loads. In this lecture, we will examine the process
of constructing influence lines for floor girders in buildings.

Consider the two-story house shown in Figure 6.233. On the main floor, the structural skeleton
of the house consists of three long girders (see Figure 6.234).

Figure 6.233: A two-story house

Consider the near exterior girder shown in Figure 6.234. It supports part of the floor. Note how
the floor rests directly on a series of short transverse beams. These beams are responsible for
transferring the floor load to the girder.

Figure 6.234: Three long girders supporting the main floor of a house

The floor is subjected to live loads (for example, furniture), which can change location on the
floor just like a vehicle can change location on a bridge. For structural analysis, when a moving

Chapter 6: Influence Lines Page 6.105


load is present, we can use influence lines to determine the load patterns that induce the
maximum internal forces in the system.

Figure 6.235 shows a moving load (red arrow) acting on the part of the floor supported by the
exterior girder.

Figure 6.235: A moving load acting on a part of the floor supported by an exterior girder

Note that the moving load is not acting directly on the girder. Rather, the load is transmitted to
the girder at the points where the transverse beams are attached to it. Our task is to construct
the influence line for the girder as the unit load moves across the floor.

A two-dimensional view of the girder, the transverse beams, the floor, and the load is shown in
Figure 6.236.

Figure 6.236: A moving load acting on a part of the floor supported by a girder

The moving load, applied directly to a segment of the floor, transfers to the girder via the
transverse beams supporting the segment. This is depicted in Figure 6.237.

Figure 6.237: A moving load transferred to a floor girder via two adjacent cross beams

The load pattern shown in Figure 6.237 is different from the pattern caused by a moving load
that acts directly on the girder. Therefore, the qualitative approach for drawing influence lines
discussed in the previous lectures is not applicable here. We need to adopt a different approach
for drawing influence lines for the floor girder.

Chapter 6: Influence Lines Page 6.106


For this analysis, we will treat the girder as a simply supported beam, resting on a pin support
and a roller support as shown in Figure 6.238.

Figure 6.238: A simply supported floor girder with seven load transfer points

When the unit load is located at the left end of the floor, acting on the left-most transverse beam,
the entire load transfers to the girder at Point A, as shown in Figure 6.239.

Figure 6.239: A unit load transferred from the floor to the girder via an exterior transverse beam

When the load moves over to the next transverse beam, again, the entire load transfers to the
girder at Point B (see Figure 6.240).

Figure 6.240: A unit load transferred from the floor to the girder via the first interior transverse beam

We can postulate that when the floor load is somewhere between these two transverse beams,
the load on the girder becomes a linear combination of the two loading cases presented in Figure
6.239 and Figure 6.240.

More specifically, we see that when the unit load is at the left end of the floor segment, Point A
on the girder would be subjected to the entire load. Similarly, we also see that when the unit load
is at the right end of the floor segment, Point B on the girder would be subjected to the entire
load. But when the unit load is in between the two segment ends, a part of the load transfers to
Point A and the rest of it transfers to Point B. This distribution of the load between the two
points is linear. Therefore, we can use the principle of superposition to determine the effect of
the load on the girder without explicitly considering all the in-between cases.

Chapter 6: Influence Lines Page 6.107


Let’s start our analysis of the floor girder by placing a unit load at each load transfer point, and
then drawing the shear and moment diagrams for the girder due to that load.

Figure 6.241 shows the shear and moment diagrams for the girder when the unit load is at Point
A.

Figure 6.241: Shear and moment diagrams for the girder due to a unit load at Point A

When the unit load is at Point C, the shear and moment diagrams for the girder can be drawn as
shown in Figure 6.242.

Figure 6.242: Shear and moment diagrams for the girder due to a unit load at Point C

Chapter 6: Influence Lines Page 6.108


When the unit load is at Point D, the diagrams shown in Figure 6.243 result.

Figure 6.243: Shear and moment diagrams for the girder due to a unit load at Point D

When the unit load is at Point E, the shear and moment diagrams for the girder are given in
Figure 6.244.

Figure 6.244: Shear and moment diagrams for the girder due to a unit load at Point E

Chapter 6: Influence Lines Page 6.109


And when the unit load is at Point G, the shear and moment diagrams become as shown in
Figure 6.245.

Figure 6.245: Shear and moment diagrams for the girder due to a unit load at Point G

Note that when the unit load is at Point B or Point F, the shear and moment in the girder are
zero since these are the support points. Figure 6.246 shows the shear and moment diagrams for
the unit load at Point B.

Figure 6.246: Shear and moment diagrams for the girder due to a unit load at Point B

Chapter 6: Influence Lines Page 6.110


FiguresFigure 6.247 and
Figure 6.248 show all of the constructed shear and moment diagrams for the girder due to a
moving unit load.

Figure 6.247: The shear diagrams for a floor girder due to a unit load

Figure 6.248: The moment diagrams for a floor girder due to a unit load

Chapter 6: Influence Lines Page 6.111


Collectively, the constructed shear and moment diagrams enable us to quickly construct shear
and moment influence lines for any point on the girder.

Suppose we wish to construct the shear influence line for a point in Segment CD of the girder.
Since the segment is not subjected to any external load between its end points, the shear force in
the segment remains constant. To draw the shear influence line for the segment, we start by
highlighting the shear values in the segment, as shown in Figure 6.249.

Figure 6.249: Highlighted shear values in a specific segment of the girder

We note that when the unit load is at Point A, the shear value in Segment CD is positive 0.5.
When the unit load is at Point C, the shear value in the segment becomes negative 0.25. When
the unit load is at Point D, the shear value in the segment is positive 0.5. When the unit load is at
Point E, the shear force becomes positive 0.25. And, when the unit load is at Point G, the shear
force is negative 0.25. These values are tabulated in Table 6.16.

Shear Force
Unit Load Location
Magnitude
A +0.5
B 0
C −0.25
D +0.5
E +0.25
F 0
G −0.25

Chapter 6: Influence Lines Page 6.112


Table 6.16: The shear force values in a girder segment due to a moving unit load

If we plot the data points given in Table 6.16 and connect them using straight lines, we get the
shear influence line for Segment CD (see

Figure 6.250).

Figure 6.250: Shear influence line for a segment of a floor girder

Suppose we also wish to draw the influence line for the shear force in Segment FG. Given the
shear diagrams in Figure 6.247, we can highlight the shear forces in Segment FG as shown in
Figure 6.251.

Chapter 6: Influence Lines Page 6.113


Figure 6.251: Highlighted shear values in a specific segment of the girder

We can then tabulate the highlighted values as shown in Table 6.17.

Shear Force
Unit Load Location
Magnitude
A 0
B 0
C 0
D 0
E 0
F 0
G +1.0

Table 6.17: The shear force values in a girder segment due to a moving unit load

The graph of the data given in Table 6.17 yields the shear influence line for Segment FG shown
in Figure 6.252.

Chapter 6: Influence Lines Page 6.114


Figure 6.252: Shear influence line for a segment of a floor girder

We can construct moment influence lines in a similar manner. Suppose we wish to draw the
moment influence line for Point B.

Upon examining the moment diagrams, we can see that only one non-zero moment is at Point B.
That is, a bending moment develops at Point B when the unit load is located at Point A. For all
other load positions, the bending moment at Point B is zero (see Figure 6.253).

Figure 6.253: Highlighted bending moment values at a specific point on the girder

Chapter 6: Influence Lines Page 6.115


The highlighted values are tabulated in Table 6.18.

Bending Moment
Unit Load Location
Magnitude
A −6.00
B 0
C 0
D 0
E 0
F 0
G 0

Table 6.18: The bending moment values at a point on a girder due to a moving unit load

If we graph the data points given in Table 6.18, we get the diagram shown in Figure 6.254. This
is the bending moment influence line for Point B.

Figure 6.254: Moment influence line for a point on a floor girder

Let’s wrap up this lecture by drawing the moment influence line for another point on the girder,
Point D. Note the moment values at Point D in Figure 6.255.

Chapter 6: Influence Lines Page 6.116


Figure 6.255: Bending moment values at a specific point on the girder

To tabulate the moment values from Figure 6.255, we can use basic geometry to determine the
height of the moment diagram when the unit load is at points A, C, E, and G. For example, when
the unit load is at Point A, the two right triangles shown in Figure 6.256 can be used to
determine the height of the moment diagram.

Figure 6.256: Two similar triangles extracted from the moment diagram for a floor girder

Equating the height-to-base ratios of the two triangles, we get 6⁄12 = ℎ⁄6. Solving the equation
for ℎ, we get ℎ = 3.0.

Chapter 6: Influence Lines Page 6.117


Similarly, when the unit load is at Point C, the height of the moment diagram can be determined
as shown in Figure 6.257.

Figure 6.257: The height of a moment diagram for a floor girder subjected to a unit load at Point C

When the unit load is at Point E, the bending moment value at Point D can be determined as
shown in Figure 6.258.

Figure 6.258: The height of a moment diagram for a floor girder subjected to a unit load at Point E

Similarly, when the unit load is at Point G, the bending moment value at Point D can be
determined as shown in Figure 6.259.

Figure 6.259: The height of a moment diagram for a floor girder subjected to a unit load at Point G

Table 6.19 shows the moment values at Point D for the various unit load positions.

Unit Load Location Bending Moment Magnitude


A −3.0
B 0
C +1.5
D +3.0
E +1.5
F 0
G −1.5

Table 6.19: The bending moment values at a point on a girder due to a moving unit load

Chapter 6: Influence Lines Page 6.118


Using the tabulated data, the moment influence line for Point D can be drawn as shown in
Figure 6.260.

Figure 6.260: Moment influence line for a point on a floor girder

Knowing the influence lines for the critical points on the girder, we can easily determine the live
load patterns on the floor that produce the maximum internal forces in the member.

For example, if the floor is expected to carry a live load of 12 kN/m, then we can use our
influence lines to calculate the maximum positive and negative moment and shear values in the
girder.

The moment influence line for Point D (see Figure 6.260) shows that the bending moment at
Point D reaches its maximum positive value when the floor is loaded as shown in Figure 6.261.

Positive area

Figure 6.261: The load pattern for the maximum positive bending moment in a floor girder

Chapter 6: Influence Lines Page 6.119


The moment value equals the area under the positive region of the diagram times the distributed
load magnitude. The positive area of the influence line is (12 × 3)⁄2 = 18. Therefore, the
maximum positive bending moment at Point D equals (18)(12) = 216 kN. m.

We can determine the maximum negative moment at Point D in a similar manner. The influence
line for the point (Figure 6.260) consists of two negative areas. Therefore, the bending moment
at Point D reaches its maximum negative value when the girder is loaded as shown in Figure
6.262.

Figure 6.262: The load pattern for the maximum negative bending moment in a floor girder

The total area of the negative regions of the influence line is (6 × 3)⁄2 + (3 × 1.5)⁄2 = 11.25.
Therefore, the maximum negative moment becomes (11.25)(12) = 135 kN. m.

We can determine other critical shear and moment values in the girder in a similar manner.

Chapter 6: Influence Lines Page 6.120


Exercise Problems:

A) Draw the following influence lines for the cantilever beam.


1) Bending moment at Point A.
2) Bending moment at Point B.
3) Shear force at Point B.

B) Draw the following influence lines for the floor girder supporting the floor.
1) Bending moment at Point A.
2) Bending moment at Point C.
3) Shear force at the hinge.

Chapter 6: Influence Lines Page 6.121


Structural Analysis I Influence Line: Floor Girder Exercise Problem Solutions

Problem 1: Draw the following influence lines for the cantilever beam shown below. The beam is a part
of a floor system.

1) Moment influence line for the fixed support (at A).


2) Moment influence line for the midpoint of the beam (at B).
3) Shear influence line for the midpoint of the beam (at B).

Solution

The floor transfers its load to the cantilever beam at two points: B and C. Therefore, we start by drawing
the shear and moment diagrams for the beam for two load cases: when a unit load is at B, and when a
unit load is at C. We also know that when the unit load is at either A or D, shear and moment in the
cantilever beam are zero.

Below is the shear and moment diagrams when the cantilever beam when subjected to a unit load at B.

Dr. Structure P a g e |1
Structural Analysis I Influence Line: Floor Girder Exercise Problem Solutions

And, when the unit load is at C, the following shear and moment diagrams result.

Therefore, the beam’s shear diagrams for the four load positions (A, B, C, and D) can be shown as
follows.

Dr. Structure P a g e |2
Structural Analysis I Influence Line: Floor Girder Exercise Problem Solutions

And, the moment diagrams for the same four load positions are:

Now we are ready to construct our influence lines. For bending moment at A, from the above diagram,
we can generate the following table.

Unit Load Position Bending Moment at A


A 0
B -4
C -8
D 0

The graph of the above data points yields the influence line for bending moment at A (see below).

Dr. Structure P a g e |3
Structural Analysis I Influence Line: Floor Girder Exercise Problem Solutions

For bending moment at B, the following table can be generated.

Unit Load Position Bending Moment at B


A 0
B 0
C -4
D 0

The graph of the above data points yields the following influence line.

Dr. Structure P a g e |4
Structural Analysis I Influence Line: Floor Girder Exercise Problem Solutions

For shear force at B, the following data points can be obtained from the shear diagrams.

Unit Load Position Shear Force at B


A 0
B +1
C +1
D 0

The graph of the above data points yields the following shear influence line.

Dr. Structure P a g e |5
Structural Analysis I Influence Line: Floor Girder Exercise Problem Solutions

Problem 2: Draw the following influence lines for the girder supporting the floor.

1) Moment influence line for the fixed support (at A).


2) Influence line for moment at C.
3) Influence line for shear at the hinge.

Solution

The floor transfers its load to the girder at two points: B and C. Therefore, we start by drawing the shear
and moment diagrams for the beam for two load cases: when a unit load is at B, and when a unit load is
at C. We also know that when the unit load is at either A or D, shear and moment in the girder are zero.

Below is the shear and moment diagrams when the girder when subjected to a unit load at B.

Dr. Structure P a g e |6
Structural Analysis I Influence Line: Floor Girder Exercise Problem Solutions

And, when the unit load is at C, the following shear and moment diagrams result.

Therefore, the girder’s shear diagrams for the four load positions (A, B, C, and D) can be drawn as shown
below.

Dr. Structure P a g e |7
Structural Analysis I Influence Line: Floor Girder Exercise Problem Solutions

And the moment diagrams for the same four load positions are:

Now we are ready to construct the influence lines. For bending moment at A, from the above diagram,
we can generate the following table.

Unit Load Position Bending Moment at A


A 0
B -3
C -3
D 0

The graph of the above data points yields the influence line for bending moment at A (see below).

Dr. Structure P a g e |8
Structural Analysis I Influence Line: Floor Girder Exercise Problem Solutions

For bending moment at C, the following table can be generated.

Unit Load Position Bending Moment at C


A 0
B 0
C +1.5
D 0

The graph of the above data points yields the following influence line.

Dr. Structure P a g e |9
Structural Analysis I Influence Line: Floor Girder Exercise Problem Solutions

For shear force at the hinge, the following data points can be obtained from the shear diagrams.

Unit Load Position Shear Force at B


A 0
B 0
C +0.5
D 0

The graph of the above data points yields the following shear influence line.

Dr. Structure P a g e | 10

You might also like