Download as pdf or txt
Download as pdf or txt
You are on page 1of 82

STRUCTURAL ANALYSIS FOR ARCHITECTS

1. INTRODUCTION
1.1 SI UNITS
The units used in the field of the strength of materials and structures are those of the Système
Internationale d’Unites; this is usually referred to as the SI system. The following are the basic
units of the SI system:
 Length : metre (m) Discuss also area, volume and derivations
Linear distances are expressed in metres and multiples or divisions of 103 of metres, i.e.
 Kilometre (km) 103m
 metre (m) l m
 millimetre (mm) 10-3 m

 Mass: kilogramme (kg) Discuss also weight


 Time: second (s)
 Temperature: kelvin (K)
- For temperatures, we use conventional degrees centigrade (“C), since we shall be
concerned with temperature changes rather than absolute temperatures.
The units which we derive from the basic SI units, and which are relevant to out
field of study, are:
 Force: newton (N) kg.m.s-2
 Work, energy: joule (J) kg.m2.s-2= Nm
 Pressure: Pascal (Pa) N.m-2= 10-5bar
The acceleration due to gravity is taken as g = 9.81m.s-2

1.2 VECTORS AND SCALARS


A scalar is a quantity that has magnitude but no direction, such as a mass, length and time. A
vector is a quantity that has magnitude and direction, such as weight, force, velocity and
acceleration

1
1.3 NEWTON’S LAWS OF MOTION
1. Newton’s First Law: Inertia
An object at rest remains at rest, and an object in motion remains in motion at constant speed and
in a straight line unless acted on by an unbalanced force.
This tendency to resist changes in a state of motion is inertia. If all the external forces cancel
each other out, then there is no net force acting on the object. If there is no net force acting on
the object, then the object will maintain a constant velocity. If that velocity is zero, then the
object remains at rest. If an external force acts on an object, the velocity will change because of
the force.
An everyday example of inertia is the motion of a ball falling down through the atmosphere and
Insert graph
2. Newtons Second Law: Force
The change of motion of an object is proportional to the force impressed; and is made in the
direction of the straight line in which the force is impressed
This has been simplified as; “The acceleration of an object depends on the mass of the object and
the amount of force applied.”
It defines a force to be equal to change in momentum (mass times velocity) per change in
time. Momentum is defined to be the mass m of an object times its velocity V.

2
Let us assume an object at a point “0” defined by its location X0 and time t0. The object has a
mass m0 and travels at velocity V0. An external force F to the object shown above moves it to
point “1”. The object’s new location is X1 at time t1. The mass and velocity of the object change
during the flight to values m1 and V1.
From Newton’s second law; 𝐹 = (m1 × V1 − m0 × V0 )/(t1 − t 2 )
Assuming the object’s mass m stays constant, the above equation becomes;
𝐹 = m ∗ (V1 − V0 )/(t1 − t 2 )
From where; F = m × a and acceleration “a” is the rate of change of velocity
From the above equations, we note that the amount of the acceleration is proportional to the size
of the force. The amount of acceleration is also inversely proportional to the mass of the object;
for equal forces, a heavier object will experience less acceleration than a lighter object.
Considering the momentum equation, a force causes a change in velocity; and likewise, a change
in velocity generates a force. The equation works both ways.
3. Newton’s Third Law: Action & Reaction
The third law states that for every action (force) in nature there is an equal and
opposite reaction.

1.4 STATIC EQUILIBRIUM


A body is said to be in equilibrium if the resultant of all forces acting on it is zero. There are two
major types of static equilibrium, namely, translational equilibrium and rotational equilibrium.
Consider a case where a book is lying on a frictionless table surface. If a perpendicular force F
is applied to the book and the book stays in the same position, in this case the vector sum of all
the force s acting on the book is zero.
∑𝑛𝑖=1 𝐹 = 𝑚𝑎 = 0
The book in this case is said to be in a state of static equilibrium i.e.
1. upward forces = downward forces
2. forces to the left = forces to the right
3. clockwise couples = counter-clockwise couples.
The equations of equilibrium for a body in a 3- dimensional system i.e. three orthogonal
directions x, y, z (Cartesian coordinate axes) are as follows;
a) ∑ Fx = 0; ∑ Fy = 0 ; ∑ Fz = 0
b) ∑ M x = 0 ; ∑ M y = 0; ∑ M z = 0

3
For planar objects lying in xy – plane with forces acting in x and y directions. The only external
moment that could act on the objet would be the one about the z-axis. The equations of
equilibrium therefore become:
∑ Fx = 0; ∑ Fy = 0; ∑ M z = 0

2. FORCES
2.1 Introduction
Definition
A force is the cause of change in the state of motion of a particle or body. It is also the product of
mass of the particle and its acceleration.
F=m×a
Force is the manifestation of action of one particle on the other. It is a vector quantity.
The derived unit for force is Newton. Discuss also KN, Relationship between kg and N
2.2 Characteristics of a force
A Force has following basic characteristics
i. Magnitude
ii. Direction
iii. Point of application
iv. Line of action
Force As A Vector Quantity
Force is a vector .i.e. an arrow with its magnitude e.g. for the force shown in Fig. 2.1, magnitude
of force is 4KN, direction is 40° with the horizontal in fourth quadrant, point of application is C
and line of action is AB.

Fig.2.1 Characteristics of a force

4
2.3 Force Systems
When a problem or system has more than one force acting, it is known as a ‘force system’ or
‘system of force’.

Fig.2.2: Force System

2.3.1 Collinear Force System


When the lines of action of all the forces of a system act along the same line, this force system is
called collinear force system.

Fig.2.3 Force System

2.3.2 Parallel Forces

Fig.2.4: Force System

5
2.3.3 Coplanar Force System
When the lines of action of a set of forces lie in a single plane is called coplanar force system.
2.3.4 Non-Coplanar Force System
When the line of action of all the forces do not lie in one plane, is called Non-coplanar force
system

Fig.2.5: Force System

2.3.5 Concurrent Force System


The forces when extended pass through a single point and the point is called point of
concurrency. The lines of actions of all forces meet at the point of concurrency. Concurrent
forces may or may not be coplanar.
2.3.6 Non-concurrent Force System
When the forces of a system do not meet at a common point of concurrency, this type of force
system is called non-concurrent force system. Parallel forces are the example of this type of force
system. Non-concurrent forces may be coplanar or non-coplanar.
2.3.7 Coplanar and concurrent force system
A force system in which all the forces lie in a single plane and meet at one point, For example,
forces acting at a joint of a roof truss (see fig.2.6)
P = External force
F1 to F5 = Member forces (internal) RA and RB = Reactions
C = Point of concurrency

6
Fig.2.6: Coplanar concurrent force system

2.3.8 Coplanar and non-concurrent force system


These forces do not meet at a common point; however, they lie in a single plane, for example,
forces acting on a beam as shown in Fig.2.7:

Fig.2.7: Coplanar non-concurrent force system

2.3.9 Non-coplanar and concurrent force system


In this system, the forces lie in different planes but pass through a single point. Example is forces
acting at the top end of an electrical pole (see Fig.2.8)

Fig.2.8: Force System

7
2.3.10: Non-coplanar and non-concurrent force system
The forces that do not lie in a single plane and do not pass through a single point are known as
non-coplanar and non-concurrent forces. Example is the loads transferred through columns to the
rectangular mat foundation as shown in Fig.2.10.

Fig. 2.10: Non-coplanar non-concurrent force system

2.4 Resolution of forces


2.4.1 Resultant and components
a) Resultant of forces: This is the sum of different forces acting together at the same
point.
b) Components of a force: A single force can be broken down into two forces which,
taken together have the same effect as the original single force
These two forces are known as components.

8
2.4.2 Some trigonometrical definitions
For any force F at an angle θ to the horizontal, the horizontal component, FX is always F.cos θ
and the vertical component, FY is always F.sin θ

(a) (b)
Fig. 2.11: Components of forces
With reference to Figure 2.11 (a),
𝐹v = 𝐹Y = F sin Ɵ
𝐹H = 𝐹X = F cos Ɵ

With reference to Figure 2.11 (b),


bc ab bc
sin Ɵ = ; cos Ɵ = ; tan Ɵ =
ac ac ab

For a triangle without a right angle in it, as seen in Figure 1.2 below, the sine and cosine rules can
be used to determine the lengths of unknown sides or the value of unknown angles.

9
The sine rule states that:

a b c
= sin 𝐵 = sin 𝐶 : where a = length of side BC; opposite the angle A
sin 𝐴
b = length of side AC; opposite the angle B
c = length of side AB; opposite the angle C

The cosine rule states that:


a2 = b2+ c2 -2bc cos A

10
WORKED OUT EXAMPLES
1. Resolution of forces
Example 1
For each of the examples given below, calculate the magnitude
and direction of the two components of the force given. Indicate the direction of the
components.

11
Example 2
Determine the forces in the plane pin-jointed framework shown below

Assume all unknown forces in each member are in tension, i.e. the internal force in each
member is pulling away from its nearest joint, as shown below.

Isolate joint A and consider equilibrium around the joint

12
1) Resolving forces vertically
upward forces = downward forces
0 = 5 + F2 cos 30
5
F2 = − cos 30 = −5.77kN (Compression)
The negative sign for F2 indicates that this member is in compression.

2) Resolving forces horizontally


forces to the left = forces to the right
F1 + F2 sin 30 = 0
F1 = −F2 sin 30 = −(−5.77) sin 30 = 2.887kN (Tension)
The positive sign for F1 indicates that this member is in tension.

The force diagram is as follows:

13
Example 3
The tension in the guy wires OA and OB of the electrical pole are 500 N and 300 N respectively
as shown in Fig.4.1. Determine the horizontal and vertical components of these tensions exerted
by the guy wires on the pole at O.

The components of each of the forces are determined as given in the following table:

Cable Force Inclination with x-axis x-component y-component


P Ө Px = P cos Ө Py = P sin Ө

OB 500 N tan-1 6/2 = 71.57° 500 cos 71.57° 500 sin 71.57°
= 158.07 N (→) = 474.36 N (↓)

OA 300 N tan-1 6/1.5 = 75.96° 300 cos 75.96° 300 sin 75.96°
= 72.78 N (←) = 291.04 N (↓)

TAKE HOME QUESTIONS


1. Determine the forces in the pin-jointed trusses shown

14
4. MOMENTS AND COUPLES
A moment is a turning effect produced by a force, on the body, on which it acts. The moment is
equal to the product of the force and the perpendicular distance of the point, about which the
moment is required and the line of action of force
A couple can be described as the moment produced by two equal and opposite forces acting
together, as shown in Figure 1.4 where,
the moment at the couple = M = F x 1(N.m)
F = force (N)
I = lever length (m)

15
Fig. 3.1: A clockwise couple
For the counter-clockwise couple of Figure 3.2,
𝑀 = 𝐹 cos Ɵ × l
where 𝐹 cos Ɵ = the force acting perpendicularly to the lever of length l.
The components of force F sin Ɵ will simply place the lever in tension, and will not cause
a moment.

Fig. 3.2: A counter-clockwise couple

16
WORKED OUT EXAMPLES
Example 1
Determine the net moment about point A for the lever systems below.

17
TAKE HOME QUESTIONS
1. Determine the net moment about point A in the systems below

18
5. DIFFERENT TYPES OF SUPPORTS
Support in a structure belongs to a member that facilitates the others member to withstand loads.
Supports in a structure transmit the load to the ground and make the structure supported on it
durable.
5.1 Fixed supports
A fixed support is the most rigid kind of support. It prevents the structural elements from
translations and rotations in all directions. Fixed supports can resist vertical and horizontal forces
as well as a moment.
The easiest example of a fixed support would be a pole or column in concrete. The pole cannot
twist, rotate or displace; it is restricted in all its movements at this connection.
Application: Fixed supports are extremely beneficial when you can only use a single support.
The fixed support provides all the constraints necessary to ensure the structure is static. It is most
widely used as the only support for a cantilever.
5.2 Pinned supports
Pinned (or hinged) supports permit rotation to happen but restrains any translation. This means it
resists vertical and horizontal forces but not moments. An easy example of a hinge is the door
that rotates only about its vertical axis without translation in any direction.
Application: Pinned supports can be used in trusses

Fig. 3.2: A Hinged support

19
5.3 Roller Supports
Rollers can only resist perpendicular forces. They cannot resist horizontal forces and of course,
moments. They are allowed to translate along the surface without any resistance to horizontal
forces.
Application: Rollers are popular in bridges; a bridge will usually have a roller support at one
end to consider vertical displacements and expansion resulting from temperature differences.
5.4 Simple Supports
This is basically when a structural member rests on other members. They are sort of similar to
rollers in such a way that they restrain vertical forces but not horizontal forces. An example is a
plank of wood resting on two concrete blocks. This kind of support is not commonly used in real
applications because there is a huge risk the member is going to fall off the support.

Fig. 4.1: Types of supports and their reactions

6. REACTIONS
Reactions are the upward forces that occur at elements’ supports in response to the applied forces
on these members.
The three rules of equilibrium are used to calculate reactions.
Consider the example shown in Fig below. A point load of 18KN is applied at 4m on a simply
supported beam of span 6m. Calculate the reactions RA and RB at the supports A and B
respectively.
From vertical equilibrium:
Total force upwards = Total force downwards
RA + RB = 18 kN
Of course, this doesn’t tell us the value of RA and it doesn’t tell us the value
of RB.
Total clockwise moment = Total anticlockwise moment

20
Taking moments about point A:
(18 kN × 4 m) = (RB × 6 m)

Fig. 4.1: Calculation of reactions for point loads


Therefore RB = 12 kN. Note that there is no moment due to force RA. This is because force RA
passes straight through the point (A) about which we are taking moments.
Similarly, taking moments about point B:
Total clockwise moment = Total anticlockwise moment
(RA × 6 m) = (18 kN × 2 m)
Therefore RA = 6 kN.
As a check, let’s add RA and RB together:
RA + RB = 6 + 12 = 18 kN

Calculation of reactions when uniformly distributed loads (UDLs) are present


In practice, most loads in ‘real’ buildings and other structures are uniformly distributed loads and
must be represented and calculated as such.
The Fig. below represents a UDL load of intensity or magnitude wkN/m of length x m applied on
a beam. Suppose that we want to calculate the moment of the UDL about point A.
We must first convert this UDL into a concentrated /point load W
In this case W = (w *x) KN and is applied at a distance x/2.
MW=(w *x *a) KNm

21
Fig. 4.2: Bending moment calculation for uniformly distributed load (UDL):
general case.
WORKED OUT EXAMPLES
Example 1
Calculate the end reactions for the beam shown in Fig. 4.3.

Fig. 4.2: Bending moment calculation for uniformly distributed load (UDL)
Vertical equilibrium: RA + RB = (3 kN/m × 2 m) = 6 kN
Taking moments about A: (3 kN/m × 2 m) × 1 m = RB × 4 m
Therefore: RB = 1.5 kN
Taking moments about B:(3 kN/m × 2 m) × 3 m = RA × 4 m
Therefore: RA = 4.5 kN
Check: RA + RB = 4.5 + 1.5 = 6 kN

Example 2

22
Determine the values of the reactions R, and RE,when a beam is simplysupported at its ends and
subjected to a downward force of 5 kN.

Solution
For this problem, it will be necessary to take moments. By taking moments, it is meant that the
values of the moments must be considered about a suitable position.
Suitable positions for takmg moments on this beam are A and B. This is because, if moments
are taken about A, the unknown section RA, will have no lever and hence, no moment about A,
thereby simplifying the arithmetic. Similarly,by taking moments about B, the unknown RB will
have no lever and hence, no moment about B, thereby simplifying the arithmetic.
Taking moments about B
clockwise moments = counter-clockwise moments
RA x ( 4 + 2 ) = 5 x 2
RA = 10/6
RA = 1.667 kN

Resolvingforces vertically
upward forces = downward forces
RA + RB = 5
RB = 5 - RA = 5 - 1.667
RB = 3.333kN

Example 3
Determine the values of the reactions of RA and RB for the simply-supported beam shown.

23
Solution
Taking moments about B
clockwise couples = counter-clockwise couples
RA x 4 = 3 x 6 + 1 0 x 2
RA = 38/4
RA = 9.5 kN

Resolvingforces vertically
RA + RB = 3 + 1 0
or RB = 13 - 9.5 = 3.5 Kn

TAKE HOME QUESTIONS


3. Determine the reactions RA and RB for the simply supported beams.

4. Determine the reactions RA and RB for the simply supported beams.

24
5. Determine the reactions RA and RB for the simply supported beams for a varying UDL

25
6. STRUCTURAL STABILITY AND DETERMINACY
A structure must not only be strong enough to be able to carry the loads and moments to which it
will be subjected, but it must also be stable. We can establish if a structure is stable or unstable
using the equation below.
m = 2j-3
Where the letter m represents the number of members in the frame and j represents the number
of joints (note that unconnected free ends of members are also considered as joints). The triangle
is the most basic stable structure.
Consider the examples below.

Fig. 4.2: Building up a framework

26
Fig. No M j Stable 2j-3 Is m = 2j-3
structure ?
(a) 2 3 No 3 No
(b) 3 3 Yes 3 Yes
(c) 4 4 No 5 No
(d) 5 4 Yes 5 Yes
(e) 6 5 No 7 No
(f) 7 5 Yes 7 Yes

From the table above, It can be shown that if m = 2j – 3 then the structure is stable. If that
equation does not hold, then the structure is not stable.
NOTE
1. A framework which contains exactly the correct number of members required to keep it
stable is termed a perfect frame. Frames (b), (d) and (f) in Fig.11.1 are examples.
2. A framework having less than the required number of members is
unstable and is termed a mechanism. In these cases, m < 2j – 3. Frames (a), (c) and (e)
in Fig. 11.1 are examples. In each case, one member of the frame is free to move relative
to the others.
3. A framework having more than this required number is ‘over-stable’ and contains
redundant members that could (in theory at least) be removed. Examples follow. In these
cases, m > 2j – 3. These frames are statically indeterminate (SI). This means
that the frames cannot be mathematically analyzed without resorting to advanced
structural techniques.

27
Examples
For each of the frames shown in Fig. below, use the equation m = 2j – 3 to determine whether
the frame is (a) a perfect frame (SD), (b) a mechanism (Mech) or (c) statically indeterminate
(SI). Where the frame is a mechanism, indicate the manner in which the frame could deform.
Where the frame is statically indeterminate, consider which members could be removed without
affecting the stability of the structure.

Fig. 4.2: Are these frameworks stable

28
Fig. No M j 2j-3 ( = or < Stable Stability
or >) structure ? Type
(a) 9 6 9 et (=) Yes SD
(b) 10 6 9 et (>) No SI
(c) 11 6 9 et (>) No SI
(d) 8 6 9 et (<) Yes Mech
(e) 6 5 7 et (<) Yes Mech
(f) 7 5 7 et (=) Yes SD
(g) 6 4 5 et (>) No SI
(h) 14 9 15 et (<) Yes Mech

The frames shown in Figs 11.2 (b), (c) and (g) are statically indeterminate. This means they are
over-stable and that one or more members may be removed without compromising stability.
In the case of Fig. 11.2 (b), any one member can be removed from the top part of the frame and
the structure would still be stable.
In Fig. 11.2 (c), two members could be removed without compromising stability – but the two
members to be removed should be chosen with care. A sensible choice would be to remove
one diagonal member from each of the two squares.
In Fig. 11.2 (g), any one member could be removed.
The frames shown in Figs 11.2 (d), (e) and (h) are mechanisms. This means that a part of the
frame is able to move relative to another part of the frame.
In Fig. 11.2 (d), the upper triangle is free to rotate about the frame’s central pin independently of
the lower part of the frame.
In Fig. 11.2 (e), thesquare part of the frame is free to deform, or collapse, as we shall see in a
later example

Frames on supports
In practice, frames have to be supported. We therefore need to consider the effects of supports on
the overall stability of frames.
In this case equation a becomes; m + r = 2j; the letter r represents the total number of restraint.

NOTE
1. If m + r = 2j, then the frame is a perfect frame and is statically determinate (SD).
2. If m + r < 2j, then the frame is a mechanism – it is unstable and should
not be used as a structure.
3. If m + r > 2j, then the frame contains redundant members and is statically
indeterminate (SI), which means it cannot be analysed without resorting to
advanced methods of structural analysis.

29
TAKE HOME QUESTIONS
For each of the frames shown in Fig. 11.5, use the equation m + r = 2j to determine whether the
frame is (a) statically determinate, (b) a mechanism or (c) statically indeterminate. Where the
frame is a mechanism, indicate the manner in which the frame could deform. Where the frame is
statically determinate, consider which members could be removed without affecting the stability
of the structure.

Fig. 4.2: Are these structures stable

4. ANALYSIS OF PIN-JOINTED FRAMES


When a beam is loaded from above, it sags resulting in deflection. The top fibres of the beam are
under compression while the bottom fibres are in tension. Deflection is the amount of downward
movement. Deflection of the beam depends on the material used, the shape and size of the
beam’s cross-section. The shallower the beam, the more the deflection, therefore in order to limit
deflection, we deepen the beam but this is not an economical solution.

30
An economical solution consists of having a framework of members of which the top and bottom
members will be in tension and compression respectively as would be in a solid beam. Such a
framework is referred to as a lattice girder or truss - it is usually made of steel but can similarly
be made of timber.
Such frameworks are analyzed as pin- jointed frames i.e. the nodes, or joints, between members
are regarded as pins, or hinges, which, by definition, cannot transmit moments from one member
to another.

Fig. 4.2: A steel railway bridge.

Fig. 4.2: Trussed bridge across River Spree, Berlin.


31
Fig. 4.2: Truss in façade, Sony Centre, Berlin.

Fig. 4.2: Roof structure, Manchester Victoria station

32
5. TRUSSES
A truss is a structure composed of slender members joined together at their end points. The
members commonly used in construction consist of wooden struts, metal bars, angles, or
channels. The joint connections are usually formed by bolting or welding the ends of the
members to a common plate, called a gusset plate.

ROOF TRUSSES

The roof load is transmitted to the truss at the joints by means of a series of purlins. The roof
truss along with its supporting columns is termed a bent. Ordinarily, roof trusses are supported
either by columns of wood, steel, reinforced concrete, or by masonry walls. To keep the bent
rigid, and thereby capable of resisting horizontal wind forces, knee braces are sometimes used at
the supporting columns. The space between adjacent bents is called a bay.
Bays are economically spaced at about 4.6 m for spans around 18 m and about 6.1 m for spans of
30 m. Bays are often tied together using diagonal bracing in order to maintain the rigidity of the
building’s structure.
Trusses used to support roofs are selected on the basis of the span, the slope, and the roof
material.

33
TYPES OF ROOF TRUSSES
a) Scissors truss – can be used for short spans that require overhead clearance.

b) Howe truss

c) Pratt truss

The Howe and Pratt trusses are used for roofs of moderate span, about 18 m to 30 m.

34
d) Fan truss

e) Fink truss

If larger spans are required to support the roof, the fan truss or Fink truss may be used. These
trusses may be built with a cambered bottom cord

35
f) Cambered fink truss

g) Warren truss

If a flat roof or nearly flat roof is to be selected, the Warren truss is often used. Also, the Howe
and Pratt trusses may be modified for flat roofs.
h) Saw tooth truss

Sawtooth trusses are often used where column spacing is not objectionable and uniform lighting
is important

36
i) Bowstring truss

The bowstring truss is sometimes selected for garages and small airplane hangars
j) Three hinged arch

Although relatively expensive, can be used for high rises and long spans such as field houses,
gymnasiums, etc

37
Assumptions for Design.
To design both the members and the connections of a truss, it is first necessary to determine the
force developed in each member when the truss is subjected to a given loading. In this regard,
two important assumptions are made in order to idealize the truss.
1. The members are joined together by smooth pins. In cases where bolted or welded
joint connections are used, this assumption is generally satisfactory provided the
centerlines of the joining members are concurrent at a point. It should be realized,
however, that the actual connections do give some rigidity to the joint and this in turn
introduces bending of the connected members when the truss is subjected to a load. The
bending stress developed in the members is called secondary stress, whereas the stress in
the members of the idealized truss, having pin-connected joints, is called primary stress.
2. All loadings are applied at the joints. In most situations, such as for bridge and roof
trusses, this assumption is true. Frequently in the force analysis, the weight of the
members is neglected, since the force supported by the members is large in comparison
with their weight. If the weight is to be included in the analysis, it is generally
satisfactory to apply it as a vertical force, half of its magnitude
applied at each end of the member.
Because of these two assumptions, each truss member acts as an axial force member, and
therefore the forces acting at the ends of the member must be directed along the axis of the
member. If the force tends to elongate the member, it is a tensile force (T), whereas if the force
tends to shorten the member, it is a compressive force (C). In the actual design of a truss, it is
important to state whether the force is tensile or compressive. Most often, compression members
must be made thicker than tension members, because of the buckling or sudden instability that
may occur in compression members.
NOTE
1. Compression
Consider a column supporting slabs, beams and other slabs in a building. The load, or force,
from all of these, is acting downwards at the top of the column. This load is represented by the
downward arrow at the top of the column. For the column to be in equilibrium, an equal upward
force / reaction develops at the bottom of the column. This reaction is represented by the
upward arrow at in Fig. (a). Not only must the rules of equilibrium (total force up = total force
down) apply for the column as a whole; these rules must apply at any and every point within a
stationary structure.

38
Fig. : Column under compression

At point C in Fig. (b), the downward force shown in Fig. (a) at point C must be opposed by an
upward force – also at point C. Thus there will be an upward force within the column at this
point, as represented by the upward broken arrow in Fig. (b). Now let us consider what happens
at the very bottom of the column – point D in Fig. (b). The upward force shown in
Fig. (a) at point D must be opposed by a downward force at the same point. This is represented
by the downward broken arrow in Fig. (b).
When a structural element is in compression: the arrows used to denote compression point
away from each other.
2. Tension
Consider a heavy metal block suspended from the ceiling of a room by a piece of string. The
metal block, under the effects of gravity, is pulling the string downwards, as represented by the
downward arrow. The string is thus being stretched and is therefore in tension.
For equilibrium, an equal upward force at the point where the string is fixed to the ceiling must
oppose this downward force. This opposing force is represented by an upward arrow in Fig. (a).
Let’s consider what happens at the top of the string. The upward force shown in Fig. (a) at point
E must be opposed by a downward force– also at this point. Thus there will be a downward force
within the string at this point, as represented by the downward broken arrow in Fig. (b). Now let
us consider what happens at the very bottom of the string – at the point where the metal block is

39
attached (point F). The downward force shown in Fig.(a) at point F must be opposed by an
upward force at this point. This upward force within the string at this point is represented by
the upward broken arrow in Fig. (b).

Fig.: String under tension

The arrows used to represent the internal forces of tension in a member point towards each
other.

40
ANALYSIS OF PIN – JOINTED FRAMES
By the term ‘analysis’ in the context of pin-jointed frames, we mean calculating:
(1) the force in each member;
(2) whether the force is tensile or compressive.

There are three techniques for doing so:


(1) Method of resolution at joints.
(2) Method of sections.
(3) Graphical method
METHOD OF RESOLUTION AT JOINTS
Guiding Rules
1. Force acts in the same direction as the member
The forces in any member of a pin-jointed frame are axial. In other words, the forces act along
the centerline of a member. If a member is vertical, the forces in that member must be vertical. If
a member is horizontal, the forces in that member will be horizontal. If a member is inclined at
an angle of, say, 30 degrees to the horizontal, the forces within the member will act along that
line.
2. Equilibrium applies everywhere
The basic rules of equilibrium apply at all nodes (and in all members) in a pin-jointed frame.
This means that the sum of all downward forces on the node exactly equals the sum of all
upward forces on the node. It also means that the total force to the left on the node exactly
equals the total force to the right.
3. Forces can be split into components
If a force acts at an angle (i.e. it is neither horizontal nor vertical), then that force can be resolved
into components – one horizontal and one vertical – which, taken together, have the same effect
as the original force. Remember, if a force F acts at an angle θ to the horizontal, its horizontal
component will always be F.cos θ and its vertical component will always be F.sin θ.

41
WORKED OUT EXAMPLES
Example 1
Determine the forces in members BD and AD. Deduce whether these members under
compression or in tension?

GENERAL APPROACH
1. Start with joints which don’t have diagonal members
2. The structure, all members and the joints are in equilibrium

At joint B;
Member AB:
 Total force up = Total force down, therefore member AB must experience a 30 kN upward force
at point B (to oppose the external 30 kN downward force). See Fig. (b)
 To achieve equilibrium in member AB, the 30 kN upward force must be opposed by a 30 kN
force downwards at the other end of the member. See Fig. (c)
 Member AB is in compression (arrows are pointing away from each other)
Member BD:
 Total force to the left = total force to the right, therefore the horizontal member BD must
experience a 64 kN rightward force at this point (to oppose the external 64 kN leftward force).
See Fig. (b)
 To achieve equilibrium in member BD, the 64 kN force to the right; must be opposed by a 64 kN
force to the left at the other end of the member. See Fig. (c)
 Member BD is in tension (arrows are pointing towards each other)

42
Example 2
Analyse the framework below and indicate which members are under compression or in tension

GENERAL APPROACH
1. Start with joints which don’t have diagonal members
2. The structure, all members and the joints are in equilibrium

At joint B:
Member BD:
 Total force to the left = total force to the right, therefore the horizontal member BD must
experience a 12 kN leftward force at this point (to oppose the external 12 kN rightward force).
See Fig. (b)
 To achieve equilibrium in member BD, the 12 kN force to the left; must be opposed by a 12 kN
force to the right at the other end of the member. See Fig. (b)
 Member BD is in compression (arrows are pointing away from each other)

43
Member AB:
 Total force up = Total force down, therefore the force in the vertical member AB must be zero
because there is no external vertical force to oppose at point B. See Fig. (b)

At joint H:
Member HG:
 Total force up = Total force down, therefore member HG must experience a 24 kN downward
force at point H (to oppose the external 24 kN upward force). See Fig. (b)
 To achieve equilibrium in member HG, the 24 kN downward force must be opposed by a 24 kN
upward force at the end G of the member. See Fig. (c)
 Member AB is in tension (arrows are pointing towards each other)

Member FH:
 Total force to the left = total force to the right, therefore the horizontal member FH must be zero
because there is no external vertical force to be opposed at point H. See Fig. (b)

At joint C:
Member CD:
 The force in the vertical member CD must be zero because there is no external vertical
force to oppose at point C.
Members AC and CE:
 Considering horizontal equilibrium at joint C, the forces in members AC and CE must be
equal and opposite – although we cannot obtain their values without further analysis.

44
Example 3
Analyse the framework below and indicate which members are under compression or in tension

At joint C:
Member HG:
 Total force up = Total force down,
Since there is no vertical external force at joint D, the member CD must experience a 0 kN. See
Fig. (b)
There is an external downward force of 60 kN there, which, for equilibrium, must be
counteracted by an upward force of 60 kN. However, member CD will not carry
this vertical force alone: diagonal members BC and CF also present at joint C will carry a
vertical component of force. Therefore, the The 60 kNupward force is shared between members
BC, CD and CF
 Forces in the members BD and DF must be equal.

45
Standard Cases

46
Example 4
1. Calculate the end reactions RA and RE
2. Analyze the framework below node by node and calculate the force in each member.

1. Determination of reactions
For vertical equilibrium, the total force up ↑ = the total force down ↓.
Therefore
RA + RE = 60 kN

2. Considering moments about point A


Clockwise moment about point A = Anticlockwise moment about point A
RE × 5 m = 60 kN × 3 m
RE = 60 kN × 3 m/5 m = 36 kN
RA = 60 – 36 = 24 kN

3. At joint A:
a) Member AB:
 Total force up = Total force down, therefore member AB must experience a 24 kN downward
force at point A (to oppose the external 24 kN upward force).
 To achieve equilibrium in member AB, the 24 kN downward force must be opposed by a 24 kN
upward force at the end B of the member.
 Member AB is in compression (arrows are pointing away from each other)

47
b) Member AF:
 Total force to the left = total force to the right, since there are no external forces at joint A, the
force in member AF, FAF, must be zero.

4. At joint E:
a) Member ED:
 Total force up = Total force down, therefore member ED must experience a 36 kN downward
force at point E (to oppose the external 36 kN upward force).
 To achieve equilibrium in member ED, the 36 kN downward force must be opposed by a 36 kN
upward force at the end D of the member.
 Member ED is in compression (arrows are pointing away from each other)

b) Member EF:
 Total force to the left = total force to the right, since there are no external forces at joint E, the
force in member EF, FEF, must be zero.

5. At joint B:
Resolving vertically
FBF × sin θ1 = 24 kN
θ1 = tan-1 (2/3) = 33.7°.
Therefore FBF × sin 33.7° = 24 kN
So FBF = 24/sin 33.7° = 43.3 kN
The vertical component of the force in member BF (at end B) must act downwards and to the right.
Because equilibrium must apply in members as well as joints, this means that the force in member BF at
end F must oppose the force at end B; in other words, it must act upwards and to the left. Because the
arrows in member BF point towards each other, member BF must be in tension.

Resolving horizontally
If the force in member BF (at end B) is 43.3 kN downwards and to the right, then the horizontal
component of this force is FBF cos θ1 = 43.3 × cos 33.7° = 36 kN (to the right).
For equilibrium, there must be an opposing (to the left) force of 36 kN and this must occur in member BC
(i.e. the only other member at joint B that can contain a horizontal force). So, the force in member BC (at
end B) is 36 kN to the left. This will be opposed by a force of 36 kN to the right at end C. Therefore, the
two arrows in member BC point away from each other, so member BC must be in compression.

6. At joint D:
Resolving vertically
FDF × sin θ2 = 36 kN
θ1 = tan-1 (2/2) = 45°.
Therefore FDF × sin 45° = 36 kN
So FDF = 36/sin 45° = 50.9 kN
The vertical component of the force in member DF (at end D) must act downwards and to the left.
Because equilibrium must apply in members as well as joints, this means that the force in member DF at
end F must oppose the force at end D; in other words, it must act upwards and to the right. Because the
arrows in member DF point towards each other, member DF must be in tension.

48
Resolving horizontally
If the force in member DF (at end D) is 50.9 kN downwards and to the left, then the horizontal component
of this force is FDC cos θ2 = 50.9 × cos 45° = 36 kN (to the left).
For equilibrium, there must be an opposing (to the right) force of 36 kN and this must occur in member
DC (i.e. the only other member at joint D that can contain a horizontal force). Therefore, the force in
member DC (at end D) is 36 kN to the right. This will be opposed by a force of 36 kN to the left at end C.
Therefore, the two arrows in member BC point away from each other, so member DC must be in
compression.

7. At joint C:
Resolving vertically
a) Member CF:
 Total force up = Total force down, therefore member CF must experience a 60 kN upward force
at point C (to oppose the external 60 kN downward force).
 To achieve equilibrium in member CF, the 60 kN upward force must be opposed by a 36 kN
downward force at the end F of the member.
 Member CF is in compression (arrows are pointing away from each other)
Resolving horizontally
The 36kN force in member BC (at end C) is to the right, therefore, to oppose this, the force in
member CD (at end C) must also be 36 kN, but to the left. The force at the other end of CD will
be to the right; therefore, the member is in compression.

Check: Resolving vertically at joint F


As elsewhere, the total force up at joint F should equal the total force down.
Vertical component of force in member:BF = FBF. sin θ1 = 43.3 × sin 33.7° = 24 kN ↑
Vertical component of force in member:DF = FDF. sin θ2 = 50.9 × sin 45° = 36 kN ↑
Vertical force in member:
CF = 60 kN ↓
Since 24 + 36 = 60, there is vertical equilibrium at joint F, so our earlier calculations are shown
to be correct. A further check could be carried out by considering horizontal equilibrium at joint
F.

49
TAKE HOME QUESTIONS
Analyze the frameworks below

50
TAKE HOME QUESTIONS
Analyze the frameworks below

51
METHOD OF SECTIONS
Oftentimes, we are interested in calculating the force in only one or two of the members of a pin-
jointed frame. In such cases, we use the method of sections instead of the method of resolution
at joints, which obliges us to determine the axial force in every member of a given pin-jointed
frame.
GENERAL APPROACH
1) Calculate the end reactions in the usual way.
2) Decide in which member(s) you need to determine the force.
3) Draw a cut line that cuts through the member(s) of interest. (The cut line may be vertical,
horizontal or inclined. It may be necessary to use different cut lines for different
members.)
4) From now on, consider the part of the frame on one side of the cut line
only (it doesn’t matter which side).
5) Use the rules of equilibrium to determine the (now external) forces in the members of
interest. Consider horizontal and/or vertical equilibrium and take moments about a
strategically chosen point. These external forces correspond to the internal forces that
existed in the members before they were ‘cut’.

52
Example 1
1. Calculate the forces in members CD, HD, and HG of the structure shown below.

Step 1: Calculation of reactions


From horizontal equilibrium of the whole structure:
HF = 15 kN (i.e. Total force → = Total force ←)
From vertical equilibrium:
VA+ VF = 50 + 20 = 70 kN (i.e. Total force ↑ = Total force ↓)
Taking moments about point A (i.e. Total clockwise moment = Total anticlockwise moment):
(50 kN × 6 m) + (20 kN × 9 m) = (VF × 12 m) + (15 kN × 4 m) + (15 kN × 4 m)

53
So, VF= 30 kN
Taking moments about point F:
(VA × 12 m) = (15 kN × 8 m) + (50 kN × 6 m) + (20 kN × 3 m)
So, VA= 40 kN
Step 2: The ‘cut’ section
we cut the frame along the cut line shown in Fig. (a). We discard the part of the frame that is
situated to the right of the cut line and will consider only the part to the left, as shown in Fig. (b).
The external forces FCD, FHD and FHG that will keep this frame in equilibrium, correspond to the
internal forces that exist in members CD, HD and HG (respectively) in the original pin-jointed
frame.
Equilibrium of the frame shown in Fig. (b)
Considering vertical equilibrium:
40 kN – 50 kN + (FHD × sin θ) = 0 (i.e. Total force ↑ = Total force ↓)
(FHD × sin θ) is the vertical component of the force in member HD.

tan θ = 4 m/3m = 1.333


Therefore θ = 53.10
40 kN – 50 kN + (FHD × sin 53.1) = 0
FHD = 12.5 kN (Tensile)

Taking moments about point H (i.e. Total clockwise moment = Total anticlockwise moment)
(FCD × 4 m) + (40 kN × 6 m) = (15 kN × 4 m)
FCD = –45 kN (Compressive)
(The minus sign indicates that the force acts in the opposite direction to that assumed – so it acts
to the left.)

Taking moments about point D


(i.e. Total clockwise moment = Total anticlockwise moment)
(40 kN × 9 m) = (50 kN × 3 m) + (FHG × 4 m)
FHG = 52.5 kN (Tensile)

TAKE HOME QUESTIONS


Use the method of sections to calculate the axial force and its sense (tension or compression) in
the members stated below for each of the pin-jointed frames shown in the figure below.
• Frame No. 1: CD, DE, EG and GH.
• Frame No. 2: BE and BF.
• Frame No. 3: BC, CD and DE

54
55
STRUCTURES AND THEIR COMPONENTS
6. STRUCTURES
A structure refers to a system of connected parts used to support a load.
The basic function of a structure is to transmit loads from the position of application of the load
to the point of support and thus to the foundations in the ground.
Any structure must satisfy the following criteria:
(1) Safety. A structure must carry the expected loads without collapsing as a whole and without
any part of it collapsing.
(2) Serviceability. The structure should not deform unduly under the effects of deflection,
cracking or vibration.
(3) Aesthetics (it must look nice).
(4) Economy (it must not cost more than the client can afford – and less if possible).
(5) Ease of maintenance.
(6) Durability. This means that the materials used must be resistant to corrosion, spalling (pieces
falling off), chemical attack, rot or insect attack.
(7) Fire resistance. While few materials can completely resist the effects of fire, it is important
for a building to resist fire long enough for its occupants to be safely evacuated.

7. STRUCTURAL ANALYSIS AND DIMENSIONING

7.1 LOADS/ACTIONS
A load is a force on a part of a structure.
During the design process, once the dimensional requirements for a structure have been defined,
it becomes necessary to determine the loads the structure must support.
In order to design a structure, it is therefore necessary to first specify the loads/actions that act on
it.
In general, the structural engineer works with two types of codes: general building codes and
design codes which specify the design loading for a structure according to its use, location (see
environmental loads) among other factors.
Examples of such codes are Eurocodes and British Standards.
Design codes provide detailed technical standards and are used to establish the requirements for
the actual structural design.
Actions are classified according to how they change over time, how they change with
position and according to their dynamic effect

56
Loads that act on structures can be divided into three broad categories: dead loads, live
loads, and environmental loads.

7.1 Types of loads


7.1.1 Dead/Permanent loads (G); are those that are constant in magnitude and fixed in location
throughout the lifetime of the structure. Usually, the major part of the dead load is the weight
of the structure itself. This is determined with good accuracy from the design configuration,
dimensions of the structure, and density of the material.
Example 1
Calculate the self-weight of a reinforced concrete beam of breadth 300 mm, depth 600 mm and
length 6000 mm.
Unit mass of reinforced concrete is 2400 kg m-3. Assuming that the gravitational constant is 10 m
s-2 (strictly 9.807 m s-2), the unit weight of reinforced concrete, ρ, is
ρ = 2400 × 10 = 24 000 N m-3 = 24 kN m-3
Hence, the self-weight of beam, SW, is
SW = volume × unit weight= (0.3 × 0.6 × 6)24 = 25.92 kN.

Calculate also its deflection under its own weight and under a live load of 5KN/m.

For buildings, floor fill, finish floors, and plastered ceilings are usually included as dead loads
and an allowance is made for suspended loads such as piping and lighting fixtures.
Normally, the dead load is not large compared to the design load for simple structures such as a
beam or a single-story frame; however, for multistory buildings it is important to have an
accurate accounting of all the dead loads in order to properly design the columns, especially for
the lower floors.
The symbols Gk and gk are used to denote the total and uniformly distributed characteristic dead
loads respectively.

57
7.1.2 Live/Imposed loads (Q) (Variable action) consist mainly of occupancy loads in
buildings and traffic loads on bridges. They may either be fully or partially in place or
not present at all, and may change in location. Their magnitude and distribution at any
given time are uncertain, and even their maximum intensities throughout the lifetime
of the structure are not known with precision. The minimum live loads for which the
floors and roof of a building should be designed are usually specified in the building code
that governs at the site of construction.

58
Fig. 2.11: Categories of use of structures

59
Fig. 2.11: Imposed loads on floors, balconies and stairs in buildings

7.1.3 Environmental loads


a) Wind loads (Variable action)
When structures block the flow of wind, the wind’s kinetic energy is converted into potential
energy of pressure, which causes a wind loading. The effect of wind on a structure depends upon
the density and velocity of the air, the angle of incidence of the wind, the shape and stiffness of
the structure, and the roughness of its surface.
Unlike dead and live loads, which are usually vertical in direction, wind loads act horizontally or
at a shallow angle to the horizontal and are therefore referred to as lateral loads

b) Snow loads (Variable action)


In countries that experience snow, roof loading due to snow can be quite severe, and therefore
protection against possible failure is of primary concern.Design loadings typically depend on the
building’s general shape and roof geometry, wind exposure, location, its importance, and
whether or not it is heated.

c) Earthquake Loads (Accidental action)


Definition: Accidental actions are those actions with a low probability of occurrence, which
are generally of short duration but have a significant effect, e.g. the effects of impacts, fires,
earthquakes and explosions
Earthquakes produce loadings on a structure through its interaction with the ground and its
response characteristics. These loadings result from the structure’s distortion caused by the
ground’s motion and the lateral resistance of the structure.Their magnitude depends on the
amount and type of ground accelerations and the mass and stiffness of the structure.

60
d) Hydrostatic and Soil Pressure
When structures are used to retain water,soil, or granular materials, the pressure developed by
these loadings becomes an important criterion for their design.Examples of such types
ofstructures include tanks, dams, ships, bulkheads, and retaining walls. Here
the laws of hydrostatics and soil mechanics are applied to define the intensity of the loadings on
the structure.

7.1.4 Other Natural Loads.


Several other types of live loads may also have to be considered in the design of a structure,
depending on its location or use.
These include the effect of blast, temperature changes, and differential settlement of the
foundation.

7.2 Nature of loads


As well as considering the different types of loading, we have to consider the nature of loads.
This could be one of three types:

a) Point load
Also called a concentrated load, this load that acts at a single point. An example would be a
column supported on a beam. As the contact area of the column on the beam would be small, the
load is assumed to be concentrated at a point. Point loads are expressed in units of
kN and are represented by an arrow in the direction that the load or
force acts, as shown in Fig 7.1 (a).

b) Uniformly distributed load


A uniformly distributed load (UDL) is a load that is evenly spread along a length or across an
area. For example, the loads supported by a typical beam – the beam’s own weight, the weight of
the floor slab it’s supporting and the live load supported by the floor slab – are consistent all the
way along the beam. UDLs along a beam (or any other element that is linear in nature) are
expressed in units of kN/m.
Similarly, the loads supported by a slab will be consistent across the slab and because a slab has
area rather than linear length, UDLs on a slab are expressed in units of kN/m2. There are at least
two different symbols used for UDL, as shown in Fig. 7.1 (b).

c) Uniformly varying load.


A uniformly varying load is a load that is distributed along the length of a linear element such as
a beam, but instead of the load being evenly spread (as with a UDL) it varies in a linear fashion.
A common example of this is a retaining wall. A retaining wall is designed to hold back earth,
which exerts a horizontal force on the back of the retaining wall. The horizontal force on the
retaining wall becomes greater the further down the wall you go. Thus the force will be zero at

61
the top of the retaining wall but will increase linearly to a maximum value at the bottom of the
wall – see Fig.
5.3 (c).
d) Volumique loads

(c) Uniformly varying loads

Fig. 7.1: The nature of loading

62
7.3 Load paths
The actual design begins with those elements that are subjected to the primary loads the structure
is intended to carry, and proceeds in sequence to the various supporting members until the
foundation is reached. Thus, a building floor slab would be designed first, followed by the
supporting beams, columns, and last, the foundation footings.

Fig. 7.1: Sequence of load transfer between elements of a


structure.

Fig. 7.1: Design process\

63
8. LIMIT STATES, LOAD COMBINATIONS, AND DESIGN LOADS,
The load cases associated with the design situations (permanent or transient) must be specified.
Every load case is characterized by a leading action corresponding to the leading hazard in the
hazard scenario and the accompanying actions occurring simultaneously with it.
8.1 Limit states
We distinguish between ultimate limit states (structural safety) and serviceability limit states.
8.1.1 Ultimate limit state
The ultimate limit states concern the safety of the structure and its ancillary elements plus the
safety of persons. We distinguish between four types:
– Type 1 is the overall stability of the structure (lateral buckling, uplift or buoyancy
as a rigid body).
– Type 2 is reaching the ultimate resistance of the structure or one of its parts
(failure due to rupture, excessive deformations, conversion of the structure into a
mechanism or loss of stability).
– Type 3 is reaching the ultimate resistance of the subsoil (landslide, slope failure,
ground failure).
– Type 4 is reaching the fatigue resistance of the structure or one of its parts.

8.1.1 The serviceability limit states


The serviceability limit states concern the functionality of the construction works, the comfort of
persons using the construction works, and the appearance of the construction works. The design
criteria for serviceability can refer to:
– Deformations that impair the functionality or appearance of the construction works or its
ancillary elements or cause damage to non-loadbearing components
– Vibrations that limit the functionality of the construction works or impair the comfort of
persons using the construction works
– Sealing defects that limit the functionality of the construction works or impair the comfort of
persons using the construction works
– Action effects specific to the type of construction, e. g. cracks or slip at connections, which
impair the appearance of the construction works and the durability of the structure
– Limit values for environmental impacts, e. g. obstructing the flow of groundwater.
NOTE
It is difficult or impossible to eliminate these phenomena i.e. deflection, cracking, or vibration –
the important thing is that the deflection and cracking are kept within certain acceptable limits.
Once the limits are exceeded, the integrity of the structure is compromised. Any type of vibration
must not have an adverse effect on the structure – this is particularly important in parts of
buildings containing plant or machinery.
Examples of failure in serviceability include;
a) A floor of a building that deflects or vibrates when in use ie when one walks on the floor.

64
b) A deflecting lintel above a doorway causing warping of the doorframe below it and
consequently the door itself does not open or close properly.
c) A crack occurring on the outside face of a reinforced concrete wall allowing may
penetration of rainwater or atmospheric humidity leads to corrosion of the steel
reinforcement within the concrete.

8.2 Design loads and load combinations


The design loads are obtained by multiplying the characteristic loads by the partial safety factor
for loads, γf The value for γf depends on several factors including the limit state under
consideration, i.e. ultimate or serviceability, the accuracy of predicting the load and the
particular combination of loading which will produce the worst possible effect on the structure
in terms of bending moments, shear forces and deflections.
In most of the simple structures, the worst possible combination will arise due to the maximum
dead and maximum imposed loads acting on the structure together.
The design will therefore be obtained from either of the following equations:
a) According to the Eurocodes; Design load = 1.35Gk + 1.5Q k
b) According to the British Standards; Design load = 1.4Gk + 1.6Q k

9. INTERNAL FORCES

9.1 Shear Force, Axial Force, And Bending Moment


These are the three internal forces that develop in an element acted upon by external forces.
a) Shear force:
Shear is a cutting or slicing action, which causes a beam to simply break or snap.
A shear force is a force tending to produce a shear failure at a given point in a beam.
The value of shear force at any point in a beam = the algebraic sum of all upward and downward
forces to the left of the point.
b) Bending moment
If a beam is subjected to a load, it will bend. The more load that is applied, the more the beam
will bend. The more the beam bends, the greater will be the tensile and compressive stresses
induced in the beam. Eventually, these stresses will increase beyond the stresses the material can
bear and failure will occur.
The bending moment is the magnitude of the bending effect at any point in a beam. The value
of bending moment at any point on a beam = the sum of all bending moments to the left of the
point.

65
c) Axial force
An axial force is a force that acts along the longitudinal axis of an element. It can either be
compression or tension. Mainly used for the analysis of columns

Fig. 7.1: Internal forces


NOTE:
M: Bending moment
V: Shear force
N: Axial force/Normal force
Given a body in equilibrium acted on by tensional (T) and compressional (C) forces on either
ends. A piece cut from this body is found to also be in equilibrium and being acted upon by the
same forces.

This cutting line is referred to as a section and it helps us investigate what is happening inside the
body. (See method of sections in the analysis of trusses)

66
We use the same principle to investigate and quantify the internal forces and stresses that
develop inside structural elements such as beams, columns, etc.

9.2 Deformation of structural elements


Under the influence of external forces, a beam undergoes a deformation depending on the
magnitude and direction of the forces and also the stiffness of the cross-section of the beam.
The stiffness of the cross-section depends on its dimensions eg
Sagging refers to downward deformation while Hogging refers to upward deformation. This
depends in part on the material used – it is obviously a lot easier to bend a beam made of rubber
than a beam of the same size made of timber!
Another factor that dictates the deflection of a beam is the shape and size of the beam’s cross-
section. If we consider a beam of rectangular cross-section, the shallower the beam is, the easier
it is to bend.

Consider a beam simply supported at either end and is subjected to a central point load. The
beam will tend to sag under that load, as indicated by the line in the corresponding
deformed/deflected diagram. When the beam has sagged, the fibers in the very top of the beam
will be in compression while the fibers in the bottom part of the beam will be in tension
Initial state Deformed shape

(a): Sagging (b) Hogging

67
Fig. 7.1: Deformations in beams – Initial state

68
Fig. 7.1: Deformations in beams – Deformed shapes

69
Fig. 7.1: Typical modes of failure for beams and columns

70
9.3 SHEAR FORCE AND BENDING MOMENT DIAGRAMS
General case 1 (Point loads)
Determine the shear force and bending moment at point X.

Shear force at X: VX = RA – M – N
Bending moment at X: MX = (RA × x1) – (M × x2) – (N × x3)

General case 2 (Point load at mid-span)

71
External reactions at supports A and B
Resolving forces vertically
Upward forces = downward forces
RA = RB = P /2 kN

Point A (X= 0) Point B (X= 1)


VA = RA = P /2 kN VB = RA – P + RB = P /2 -P + P/2 =0 kN
MA = 0 kNm MB = (P/2*L) -(P* L/2) = 0 kNm

Point C (X= L /2)


VC = RA – P = P /2 -P = -P/2kN
MC = (P/2*L/2) = PL/4 kNm

General case 3 (Point load at a distance a)

72
External reactions at supports A and B
Resolving forces vertically
Upward forces = downward forces
RA + RB = P kN …………………… (1)

Taking moments about A


Clockwise moments = counter-clockwise moments
(P* a) = (RB * L)
RB = Pa/L kN …………………… (2)

Taking moments about B


Clockwise moments = counter-clockwise moments
(RA * L) = (P* b)
RA = Pb/L kN …………………… (2)

Point A (X= 0) Point B (X= 1)


VA = RA = Pb/L kN VB = RA – P + RB = (Pb/L - P + Pa/L) kN
MA = 0 kNm MB = (Pb/L*L) -(P* b) = 0 kNm

Point C (X=a)
VC = RA – P = (Pb/L – P) kN
MC = (Pb/L*a) = Pab/L kNm

73
Example 1
Determine the shear force and bending moment at points A, B, C, D, E, F, and D and
draw the shear force and bending moments diagrams

External reactions at supports A and G


Resolving forces vertically
Upward forces = downward forces
RA + RG = 18kN …………………… (1)

Taking moments about A


Clockwise moments = counter-clockwise moments
(18* 4) = (RG * 6)
RG = 72/6 = 12kN …………………… (2)

74
(2) into (1)
RA = 6kN

Point A (X= 0) Point B (X= 1)


VA = 6 kN VB = 6 kN
MA = 0 MB = + (6*1) = 6 kNm

Point C (X= 2) Point D (X= 3)


VC = 6 kN VD = 6 kN
MC = + (6*2) = 12 kNm MD = + (6*3) = 18 kNm

Point E (X= 4) Point F (X= 5)


VE = 6 -18 = -12 kN VF = 6 -18 = -12 kN
ME = + (6*4) = 24 kNm MF = + (6*5) – (18*1) = 12 kNm

Point G (X= 6)
VG = 6 -18+12 = 0 kN
MG = + (6*6) – (18*2) = 0 kNm

General case 3 (Point load at a distance a)

75
External reactions at supports A and B
Resolving forces vertically
Upward forces = downward forces
RA = RB = wL/2 kN …………………… (1)

Point A (X= 0) Point B (X= L)


VA = RA = wL/2 kN VB = wL/2 – Wl + wL/2 = 0kN
MA = 0 kNm MB = (wL/2 *L) -(wL*L/2) = 0 kNm

Point C (X=L/2)
VC = (wL/2 – wL) = -wL/2 kN
MC = (wL/2 *L/2) - (wL/2*L/4) = wL2/8 kNm

Example 2
Draw the shear force and bending moments diagrams for the beam below that’s charged
by UDL of 4 KN/ml

76
External reactions at supports A and G
Resolving forces vertically
Upward forces = downward forces
RA = RG = (4 kN/m × 6m) /2 = 12 kN

Point A (X= 0) Point B (X= 1)


VA = 12 kN VB = 12- (4*1) = 8 kN
MA = 0 MB = + (12*1) -(4*1* 0.5) = 10 kNm

Point C (X= 2) Point D (X= 3)


VC = 12 – (4*2) = 4kN VD = 12 – (4*3) = 0kN
MC = + (12*2) - (4*2* 1) = 16 kNm MD = + (12*3) - (4*3* 1.5) = 18kNm

Point E (X= 4) Point F (X= 5)


VE = 12 – (4*4) = -4kN VF = 12 – (4*5) = -8kN
ME = + (12*4) - (4*4* 2) = 16 kNm MF = + (12*5) - (4*5* 2.5) = 10 kNm

Point G (X= 6)
VG = 12 – (4*6) +12 = 0kN
MG = +(12*6) - (4*6*3) = 0 kNm

77
Example 3
Draw the shear force and bending moments diagrams for the system below

External reactions at supports A, B and C


Resolving forces vertically
Upward forces = downward forces
RA, RB, RC have already been found
RA = -2.01t, RB =23.69t, RC = 12.325t
78
Point A (X= 0)
VA = RA = -2.01t
MA = 0

Point A’ (X= 2m)


VA’ = RA - 4 = -2.01 - 4 = -6.01t
MA’ = -2.01*2 = -4.04tm

Point B (X= 4m)


VB = RA - 4 + RB = -2.01 - 4 + 23.69 = 17.68t
MB = -2.01*4 – 4*2 = -16.04tm

Point B’ (X= 7m)


VB’ = RA - 4 + RB -6- 4*3 = -2.01 - 4 + 23.69 - 6 - 12 = -0.32t
MB’ = -2.01*7 – 4*5 + (23.69*3) - (4*3*1.5) = 19tm

Point C (X= 10m)


VC = RA - 4 + RB -6- 4*6 + RC = -2.01 - 4 + 23.69 - 6 – 24 + 12.325 = 0t
MC = -2.01*10 – 4*8 + (23.69*6) - (4*6*3) - 6*3 = 0tm
See further explanation on the diagram
NOTE
Alternatively make section/cut K-K through the beam, and study what is to the right
of this section.

When we resolve vertically, the general equation for the shear force at C;
T(x) + RC - 4*x = 0; From where, VC = T(x) = 4*x -RC
At x = 0; VC = -RC = -12.325t

79
Example 4
Draw the shear force and bending moments diagrams for the system below

1. External reactions at supports A and E


a) Resolving forces vertically
Upward forces = downward forces
RAV + RE = 15sin60 + 30 + 3.5*10 +20
RAV + RE = 98 kN …………………… (1)

Taking moments about A


Clockwise moments = counter-clockwise moments
(15sin60*3) + (30*6) + (3.5*10 *6.25) + (20*10) = (RE * 12)
RE = 53. 14 kN …………………… (2)

(2) into (1)


RAV = 44.85 kN

80
b) Resolving forces horizontally
Leftward forces = Rightward forces
RAh + 15cos 60 = 10 kN
RAH = 2.5 kN

2. Finding internal forces at various points along the beam

Point A (X= 0)
VA = RAV = 44.85kN
MA = 0

Point B (X= 3m)


VB = RAV – 15sin 60 = 44.85 – 15sin 60 = 37.35 kN
MB = 44.85 *3 = 134.55 kNm

Point B’ (X= 4.5m)


VB’ = RAV – 15sin 60 = 44.85 – 15sin 60 = 37.35 kN
MB’ = 44.85 *4.5 – 15sin 60* 1.5 = 182.34 kNm

Point C (X= 6m)


VC = RAV – 15sin 60 – (10* 1.5) -30 = 44.85 – 15sin 60– (10* 1.5) -30 = -13.14 kN
MC = (44.85 *6) – (15sin 60* 3) - (10* 1.5*0.75) = 218.88 kNm

Point B’’ (X= 8m)


VB’’ = RAV – 15sin 60 – (10* 3.5) -30 = 44.85 – 15sin 60– (10* 1.5) -30 = -33.14 kN
MB’’ = (44.85 *8) – (15sin 60* 5)- (10* 3.5*1.75) –(30*2) = 172.60 kNm

Point D (X= 10m)


VD = RAV – 15sin 60 – (10* 3.5) -30 -20
= 44.85 – 15sin 60– (10* 3.5) -30 -20
= -53.14 kN

MD = (44.85 *10) – (15sin 60* 7)- (10* 3.5*3.75) –(30*4) = 106.32 kNm

Point E (X= 12m)


VE = RAV – 15sin 60 – (10* 3.5) -30 -20 -53.14
= 44.85 – 15sin 60– (10* 3.5) -30 -20 -53.14
= 0 kN

ME = (44.85 *12) – (15sin 60* 9)- (10* 3.5*5.75) –(30*6) –(20*2) = 0 kNm

81
82

You might also like