You are on page 1of 18

2: Introduction to Structures and

Structural Behaviour

Summary
Unit 2 deals with the behaviour of structures in buildings. An understanding of
the types of loads and forces, and how they are handled by materials and
components is a necessary prerequisite for selecting the correct material for the
purpose used. Unit 2 uses two sources of reference. Chapter 1 of Reid (1984) is
presented as Appendix B. It is a descriptive introduction to the types and
functions of structural systems used in buildings. As you study this material
work through the ITQs. Chapter 1 of Beckmann and Bowles (2004) is
presented as Appendix C and explains the principles of structures very
concisely, with more emphasis on quantitative understanding. If you are
already familiar with this kind of material you will find the ITQs in this unit
straightforward so just use them to refresh your understanding. However, if the
material is new to you it will be necessary to work quite hard at the ITQs and
perhaps supplement your reading.

In the In-Text Questions the chapter, page and figure references are to the
chapter, page and figure numbers in the relevant Appendices (B and C).

1.1 Appendix B - Reid


Study Appendix B and work through the following ITQs.

ITQ 2.1. (p.3)


What is the difference between strength and stiffness, and between stress and
strain?

ITQ 2.2. (p.4)


Why do we stagger or overlap the individual bricks/blocks in masonry?

ITQ 2.3. (p.4)


Why is a slender wall unstable?

ITQ 2.4. (p.4)


Make a list of the types of forces that cause loads in buildings to deviate from
vertical.

ITQ 2.5. (p.5)

Heriot-Watt University Unit 2


2: Introduction to Structures and Structural Behaviour

Considering Fig. 1.9, draw a triangle of forces to scale to estimate the largest
horizontal force that the 3kg brick can resist without overturning.

ITQ 2.6. (p.5)


Why is it necessary for a tall wall to become thicker towards the base?

ITQ 2.7. (p.5)


Does adding piers or stepping a wall increase or decrease the effective
thickness?

ITQ 2.8. (p.6)


Why can a column not be built as high as a wall of the same thickness?

ITQ 2.9. (p.6)


Why did the ancients Greeks space their columns no more than 4m apart?

ITQ 2.10. (p.6)


Considering the arch (Fig. 1.15), use the triangle of forces to explain why the
arch develops an outwards thrust into the banks of the river.

ITQ 2.11. (p.7)


How does the keystone function?

ITQ 2.12. (p.7)


Why did Wren use a chain around the base of the dome of St Paul’s cathedral?

ITQ 2.13. (p.8)


Sketch the flow of compressive forces down to the ground on the cross-section
of the Gothic cathedral in Fig. 1.18.

ITQ 2.14. (p.9)


Referring to Fig.1.19, if the load increases does the maximum bending moment
increase or decrease? If the span decreases does it increase or decrease? Is the
bending moment affected by the width or depth of the beam? If so, in what
sense?

ITQ 2.15. (p.9)


Figs. 1.19d and 1.19e show the variation of bending moment along the beam
for a point load and a uniformly distributed load. If the same total load is
applied to a beam is the bending moment larger, the same or smaller when the
load is a point load or a uniformly distributed load?

ITQ 2.16. (p.9)


Considering the stress distribution shown in a bending beam in Fig. 1.20, if the
material in the beam is weaker in tension than in compression, would you

Unit 2 Heriot-Watt University


expect the beam to fail by cracking on the bottom surface or crushing on the
top surface?

ITQ 2.17. (p.9)


Figs 1.21a and 1.21b show how shear forces develop in a beam. Is the stress in
a higher, the same or lower than in b? If shear stress exceeds the shear strength
of the material it will break. Since metals have the same shear strength in every
direction whereas timber is considerably weaker in shear along the grain than
across the grain, would you expect metals and timber to be more likely to break
in shear along the beam (case a) or across (case b)?

ITQ 2.18. (p.10)


The internal moment of resistance is what enables a beam to withstand the
external applied bending moment. Is the moment of resistance increased (i.e.
the beam is stronger) or decreased (weaker) when the depth, width or span of
the beam is increased?

ITQ 2.19. (p.11)


An efficient structural member is one that contains no more material than is
needed to resist the loads on it. Therefore, why is a castellated beam more
efficient than a universal beam?

ITQ 2.20. (p.11)


Why is stiffness more important than strength in a beam or truss?

ITQ 2.21. (p.12)


In a reinforced concrete beam where are the steel bars placed in relation to the
anticipated loading pattern?

ITQ 2.22. (p.17)


Why is an encastré beam stronger than a simply supported one? (Aside:
cantilevered stairs are an example of a series of encastré beams.)

ITQ 2.23. (p.18)


Why are the knees of a portal frame thickened?

ITQ 2.24. (p.19)


In what way does the shape of a catenary mirror that of an arch? Why is this
useful in understanding how an arch works structurally?

ITQ 2.25. (p.22)


What are the main ways of giving a multi-storey frame stability against wind
forces?

ITQ 2.26. (p.24)


What is the essential purpose of a foundation?

ITQ 2.27. (p.25)

Heriot-Watt University Unit 2


2: Introduction to Structures and Structural Behaviour

What is differential settlement? What are its consequences?

ITQ 2.28. (p.26)


What is the difference between a one-way spanning floor slab and a two-way
slab?

ITQ 2.29. (p.27)


What are the advantages and disadvantages of the cantilever slab / curtain
walling system described here?

ITQ 2.30. (p.30)


What is flutter in a membrane roof and how can it be avoided?

ITQ 2.31. (p.30)


Tension structures in buildings are quite rare. Why do you think that is?

ITQ 2.32. (p.32)


The difference in load carrying capacity between a flat sheet and a folded sheet
is considerable. Think of everyday examples of structural forms, both natural
and man-made, that exploit that difference to make structurally useful items.

ITQ 2.33. (p.33-34)


Hyperbolic paraboloid surfaces became fashionable because they could be
economically cast in concrete. Why is this possible? (Aside: it was only the
application of computer aided design and manufacture in the late 20th century
that made it possible to produce “natural” forms using steel frames and
lightweight cladding – e.g. Gehry’s Guggenheim Museum, Bilbao – and these
are now commonplace.)

2.2 Forces and stresses


The forces exerted on materials and structures in buildings come from several
sources:
• The self-weight of materials within the structure and fabric;
• The weight of people and contents (furniture, equipment, goods);
• Environmental loadings such as wind and snow.
It is normal to categorise loads as either ‘dead’ - in the sense that they do not
change over the lifetime of the structure (e.g. the self-weight of the materials) -
or ‘imposed’ – which vary as people move around, and wind / snow comes and
goes. Many of these forces are weights and we can understand them as follows.
An object has a mass m, reported in kg. On Earth it experiences a force P due
to gravitational attraction. This attraction is given as g, the acceleration under
gravity, with the value 9.81 m/s2. The weight of a mass m is therefore the force
P, given by

Unit 2 Heriot-Watt University


P=m×g
Force has units of Newtons (N), so for a mass of 100 kg, the weight (force) is
981 N. On the moon, where gravity is one-sixth that on Earth, the weight
(force) exerted by a 100 kg astronaut is only 163.5 N. In everyday terms we
take the gravitational attraction for granted and speak of the weight of an object
when we mean the force it exerts under gravity. As a result force is also given
units of kilogram force (kgf), which is defined as the force exerted by a mass of
1 kg on Earth, so 1 kgf = 9.81 N

Forces have direction as well as size. Gravity always acts downwards, so


weights do too, whereas wind always acts horizontally. When thinking about
the combined action of two or more forces, we cannot simply add them
together numerically, instead we must calculate the resultant, using the so-
called triangle of forces. Fig. 2.1 shows that the resultant of two forces acting at
right angles can be represented by the line completing the right angled triangle
(the hypotenuse): the size and direction can be calculated using trigonometry or
estimated by drawing the forces to scale on squared paper and measuring the
length of the line.

Fig. 2.1 The resultant of two perpendicular forces acting at the same
point has the size and direction of the hypotenuse of the triangle.

The inverse process is also helpful in understanding forces: any force can be
resolved into a vertical and horizontal component (Fig. 2.2), the size of which
can be calculated using trigonometry or estimated by a scale drawing.

Fig. 2.2 The downward vertical component of the force P is Psinθ and
the horizontal leftward component is Pcosθ.

As described in Unit 1, forces are converted to stresses (defined as force per


unit area of application) within the materials to which they are being applied.
The three most common stress configurations are compression, tension and
shear (Fig. 2.3) and most loads in buildings can be analysed in terms of these
three stresses.

Heriot-Watt University Unit 2


2: Introduction to Structures and Structural Behaviour

Fig. 2.3 In each stress configuration the stress level is determined by


the force P divided by the cross sectional area of the plane
(compression, tension or shear) marked by the dashed lines over which
the force is applied.

Finally, in structures loads may be either point loads or uniformly distributed


(UDL). Thus, on Earth the 100 kg astronaut exerts a point load on a floor of
981 N. If the contact area of the feet is 200 × 300 mm, the compressive stress is
0.01635 N/mm2, but because that contact area is small in comparison to the
whole floor area of 3 × 5 m it can be considered to be a point load. However,
the flooring materials apply a uniformly distributed dead load to the whole
floor given by:
(volume of material × density × g) / floor area.
For example, 1 m2 of 100 mm thick wood block flooring of density 500 kg/m3
has a mass of 1 × 1 × 0.1 × 500 kg and exerts a load of 50 kgf or 490.5 N. This
corresponds to a UDL stress of 490.5 N/m2.

2.3 Moments
When the applied forces are not in line (for example as in the shear situation
sketched in Fig. 4.3) the result is a turning moment. This is the same action as a
lever and the example of a children’s seesaw (Fig. 2.4) shows the principle. A
35 kg child sitting 1.4 m from the fulcrum generates a moment of 49 kgf-m or
480.7 Nm.

Fig. 2.4 Each child on the seesaw exerts a moment given by the force
multiplied by the distance from the fulcrum.

Unit 2 Heriot-Watt University


2.4 Equilibrium
All buildings and structures are static, which means that they must be in
equilibrium, because any object that is not in equilibrium will move until it is
in equilibrium. Equilibrium means that every force and every moment must be
balanced by an equal and opposing force and moment, and this applies at every
point in the structure. In the example of the seesaw (Fig. 2.4) unless the
children are in equilibrium one of them will move up and the other down. If the
other child is 40 kg and is sitting 1.25 m from the fulcrum his moment is 50
kgf-m or 490.5 Nm. In order to achieve equilibrium, this child must move
further in to 1.225 m or alternatively the lighter child should move further out
to 1.428 m from the fulcrum.

Likewise the forces must be in equilibrium. Neglecting the weight of the beam
itself, the total downward forces on the fulcrum are 75 kgf-m or 735.75 N.
These must be balanced by the upward reaction force exerted by the fulcrum on
the beam. If the downward forces increase as more children climb aboard, so
does the reaction force increase to maintain equilibrium. Finally, both vertical
and horizontal components of every force must also be balanced to ensure
equilibrium, so when analysing forces and checking for equilibrium, it is best
to resolve all forces into their vertical and horizontal components first.

2.5 Beckmann and Bowles - Appendix C


Study Appendix C sections 1.1 to 1.2.5 and work through the following ITQs.

ITQ 2.34. (p.3)


The mass of the lantern in Fig. 1.1 is 10 kg. Confirm that the tension force W in
the hanging rod is 10 kgf or 98.1 N.

ITQ 2.35. (p.3)


The same lantern is supported by a bracket 900 mm long (Fig. 1.4). Confirm
that the anticlockwise turning moment M at the point where the cantilevered
bracket enters the wall is 9 kgf-m or 88.3 Nm.

ITQ 2.36. (p.4)


What happens to an object if the forces and/or moments acting on it are not of
equal size and acting in opposite directions?

ITQ 2.37. (p.5)


Taking the situation depicted in Fig. 1.6, the men move closer together, so that
S is now l/3 from the left hand post and B is now l/3 from the right hand post,
with l/3 between them. What happens to the two reactions RL and RR? What
happens to the bending moment in the seat at the mid-point?

ITQ 2.38. (p.8)

Heriot-Watt University Unit 2


2: Introduction to Structures and Structural Behaviour

For the situation in Fig. 1.8, the steel beam is 6 m long supporting a brickwork
wall 3 m high and 225 mm thick. By how much does the bending moment at
the mid-point from this uniformly distributed load change if the wall increases
to 4 m high? Or to 450 mm thick?

ITQ 2.39. (p.8)


Considering the steel beam in ITQ 4.38 by how much does the bending
moment at the mid-point increase if the span increases from 6 m to 7.5 m?

ITQ 2.40. (p.8)


For the situation shown in Fig. 1.9b why is there an upward net force of ⅜L at
the node above the wall in the roof truss?

ITQ 2.41. (p.9)


For the roof truss in Fig. 1.9 use a scale drawing of the triangle of forces to
confirm that the forces T and C are approximately ⅜L and 0.53L, respectively,
when the angle  = 45.

ITQ 2.42. (p.9)


In general terms what will happen to the roof truss and walls shown in Fig. 1.9
if the tie beam is cut through?

ITQ 2.43. (p.10)


The hanging rod in Fig. 1.5 is 6 mm in diameter. How does the tension stress 
in it, caused by the weight of the lantern, change if the rod is exchanged for one
of 8 mm in diameter?

ITQ 2.44. (p.11)


An historic church clock is driven by a lead casting of mass 25 kg suspended
by a multistrand steel wire 1.5 mm in diameter. After a week’s operation the
casting has reached the bottom of the church tower 10 m below the escapement
mechanism and the wire has extended by 6.6 mm. If the steel wire is
exchanged for one 2 mm in diameter how will the extension change? If the
steel wire is exchanged for an aluminium wire of the same diameter how will
the extension change? (Modulus of elasticity of steel = 210 kN/mm2 and of
aluminium = 70 kN/mm2).

ITQ 2.45. (p.11-12)


Why is ductile fracture preferable to brittle fracture in an engineering material?

ITQ 2.46. (p.13-14)


How do the tests described here show that compressive or crushing failure in a
brittle material is actually a failure in tension?

ITQ 2.47. (p.13)

Unit 2 Heriot-Watt University


To determine the apparent stress at failure, given by force P  area A,
cylindrical specimens of sandstone 150 mm diameter are tested in a
compression test like that described in Fig. 1.12. Why is the measured failing
stress of a specimen 150 mm high twice that of one 450 mm high? What would
be the effect of fixing a strong steel band around the latter at mid-height?

2.5.1. Bending behaviour of beams


Beckmann and Bowles (Appendix C) explain that at equilibrium in a beam the
externally applied bending moment must be resisted by the internally generated
moment between the tension force T and the compression force C in Fig. 1.16.
When a simply supported beam deflects under load the top half is in
compression and the bottom half is in tension. At some point in between there
is a neutral axis where the material is in neither compression nor tension and at
this point the stress is zero. Moving away from the neutral axis the stress
increases in proportion to distance, reaching a maximum at the surface.

ITQ 2.48. (p.15-16)


What happens to the load-carrying capacity of the beam if it is made deeper?

2.5.2. Shear in beams


Both Reid (Appendix B, Fig. 1.21) and Beckman and Bowles (Appendix C,
section 1.2.4) explain how shear forces are produced in beams. On the left of
the person in Fig. 1.21 the shear generated by the point load acting downwards
and the reaction at the left hand support acting upwards is ‘clockwise’, whereas
on the right of the person the shear is ‘anti-clockwise’. This important feature
means that the shear forces due to a point load are constant between load and
support but change direction in each half of the beam and are zero at the
precise point where the load is applied. In contrast, a uniformly distributed load
generates shear forces which vary up to a maximum at the supports but pass
through zero at mid-span and, again, change direction in each half of the beam,
as shown in Fig. 4.5, where the wavy line is the conventional symbol for a
uniformly distributed load.

Heriot-Watt University Unit 2


2: Introduction to Structures and Structural Behaviour

Fig. 2.5 A uniformly distributed load on a beam generates shear forces


that vary along the beam but are greatest at the supports. If the shear
stress exceeds the shear strength of the material the beam will fail at
the supports but because the shear stress is low around mid-span a
smaller cross-section there can still resist the shear. An alternative used
in reinforced concrete beams is to add steel reinforcement near the
supports arranged vertically to increase the shear resistance.

ITQ 2.49. (p.16-17)


What happens to the shear resistance of the beam if it is made deeper?

ITQ 2.50. (p.17)


Why is shear failure in a brittle material, such as concrete, undesirable?

ITQ 2.51. (p.17)


Using the triangle of forces show that the shear forces in Fig 1.18(b) are
equivalent to the inclined compression and tension forces in Fig. 1.18(c).
Hence, explain why a brittle material cracks at 45 to the shear forces.

2.5.3. Deflections in beams


Beckmann and Bowles (Appendix C) explain how all materials shorten in
response to compressive loads and lengthen in response to tensile loads: thus
the top surface of a beam supporting a point load shortens and the bottom
surface lengthens and consequently the beam bends. This deflection is
inevitable but excessive deflection is unacceptable to users. For example
people do not like walking across a springy plank bridge, they do not like
looking at beams that sag excessively and deflections may crack brittle plaster

Unit 2 Heriot-Watt University


finishes or cause windows and doors to jam instead of working freely. In
practice, deflection, not strength, is normally the limiting criterion in beam
performance. Typically, maximum deflections of between span/360 and
span/200 are allowed, depending on the situation: that corresponds to 10 mm
deflection in a span of between 2 m and 3.6 m.

The deflection depends on load, span, material stiffness (i.e. the modulus of
elasticity) and the geometry of the beam section. The higher the load the larger
the deflection will be and the greater the span the larger the deflection will be.
On the other hand, the higher the modulus of elasticity (stiffer material) the
smaller the deflection. The effect of section geometry is more subtle and is
alluded to by Reid (Appendix B), p.10-11. In the same way as a deeper beam
has a larger moment of resistance because the compression and tension zones
are spaced further apart so a deeper beam is stiffer and deflects less. In a beam
with a solid rectangular cross-section the deflection under a load is inversely
proportional to the beam’s depth cubed, so doubling the depth of a beam
reduces the deflection 8-fold. Likewise the deflection is inversely proportional
to the beam’s width, so doubling the width of the beam halves the deflection.
Structural engineers define the cross-section of the beam in terms of the second
moment of area or moment of inertia and handbooks provide formulae to
calculate the second moment of area of all kinds of solid or hollow section
beams. For the solid section beam it is (width × depth-cubed ÷ 12). The
important point is that the second moment of area takes the shape of the beam
section into account and doubling the second moment of area halves the
deflection: the two are inversely proportional.

ITQ 2.52.
If everything else remains unchanged is the deflection of a beam increased or
decreased when its depth, width or span is increased?

ITQ 2.53.
To lighten a structure a beam is changed from steel to aluminium of the same
cross-section. How will the deflection change?

ITQ 2.54.
To lighten a structure a solid beam is changed to a hollow one of the same
second moment of area. How will the deflection change?

2.5.4. Buckling of compression members


A long member acting in compression, whether it is a masonry column or a
strut in a truss, is more likely to fail by buckling than by crushing. For the
structural engineer the question is where is the point of transition between a
short, stocky member and a slender member.

The strength of a short column is based only on its cross-sectional area and the
material strength, but Reid (Appendix B) points out that a slender column fails
by sideways buckling. The buckling results in a tensile failure on the outer
surface of the curved column and common-sense indicates that the length of

Heriot-Watt University Unit 2


2: Introduction to Structures and Structural Behaviour

the column plays a significant role. For example, a matchstick is reasonably


strong in compression along its length but a 300 mm long stick of the same
cross-section would be very weak. This observation is dealt with by
introducing the concept of slenderness ratio (= effective length / effective
thickness) into structural design processes.

All design codes provide curves or tabulated values of the load-carrying


capacity in compression, showing how it decreases as slenderness ratio
increases. They are different for different materials but the principles are the
same whether we are dealing with steel struts, aluminium alloy struts or
brickwork or other masonry columns. The effective length is the actual length
in the case of a pin-jointed member, but is reduced in situations where greater
restraint against rotation is provided at the ends of the member because this
effectively stiffens the member. For example, a steel column strongly bolted
down to a foundation but pinned at the top has an effective length two-thirds of
its actual length. The effective thickness is taken as the least lateral dimension
of a solid column or given as a geometrical factor for other sections like I-
beams or tubes and tabulated in structural design handbooks.

Permissible slenderness ratios depend on the material being considered.


Masonry columns are considered slender if the slenderness ratio exceeds 8,
concrete columns if it exceeds 10, and the permissible stress has to be reduced
accordingly to prevent buckling and tensile failure. The upper limit of the
slenderness ratio in these materials is 27 provided that the load is applied
axially, but is reduced if the load is eccentrically applied. Fig. 4.6 explains how
eccentricity is calculated in a solid wall and Table 4.1 gives values of the so-
called capacity reduction factor that must be applied to the strength of the
masonry to give the permissible load.

Fig. 2.6 Eccentric loading is defined by the eccentricity, given by the


distance between the point of load application and the centre line of the
wall, divided by the overall thickness of the wall.

Unit 2 Heriot-Watt University


Table 2.1 Capacity reduction factors for masonry. The load carrying
capacity is given by the masonry strength multiplied by the relevant
value given in the table. (From BS5628).

Slenderness Eccentricity at top of wall ex


ratio 0.05 0.1 0.2 0.3
0 1.00 0.88 0.66 0.44
6 1.00 0.88 0.66 0.44
8 1.00 0.88 0.66 0.44
12 0.93 0.87 0.66 0.44
18 0.77 0.70 0.57 0.44
22 0.62 0.56 0.43 0.37
24 0.53 0.47 0.34 -
26 0.45 0.38 - -
27 0.40 0.33 - -

Buckling failure is also a consideration in beams. Recalling that an I-section


steel universal beam is efficient because the material is concentrated in the
flanges to maximise the internal moment of resistance, one might be tempted to
make the web as thin as possible. This is not a good idea because even the
relatively small compressive force in the region between the neutral axis and
the flange could be enough to cause buckling failure of a thin web. (See
Beckmann and Bowles, Appendix C, fig. 1.19.) The web thickness in standard
steel sections is prescribed in order to avoid this problem.

ITQ 2.55.
Confirm that the slenderness ratio of the masonry column of a cathedral nave
which is 750 mm in diameter and 18 m high, restrained top and bottom
(effective length = height / 2), is 12. Compare this to the values mentioned in
the text.

ITQ 2.56.
Using table 2.1, by how much would the load-carrying capacity of the column
in ITQ 2.55 change if the load became eccentric, with eccentricity changing
from 0.05 to 0.3.

ITQ 2.57.
Look again at ITQ 4.19 in the light of buckling resistance. Recognising that
the compression force in the web is proportional to the distance from the
neutral axis, what is the best shape for the cut-outs in the web of a
castellated beam?

2.6 Frame structures and trusses


Any reasonable sized building depends on the use of frames (which are
normally clad with suitable panelling) to enclose space and protect from the
weather. Frames consist of members and the joints between them and these

Heriot-Watt University Unit 2


2: Introduction to Structures and Structural Behaviour

joints may be pinned, where all of the members are free to rotate relative to
each other, or fixed, where they are not. The simplest pin jointed structure is a
triangle – three members, three joints – and in fact this is the only truly stable
one. A four- or more-jointed structure like this is completely floppy and
useless. The principle of triangulation – to make a series of three membered
units – is key to the satisfactory performance of most structural frames. The
difficulty of making sufficiently reliable fixed joints in timber buildings forced
carpenters throughout history to reinforce joints with triangular bracing, which
can typically be concealed within the wall panelling. Thus the simple floppy
frame building in Fig. 2.7 is stabilised by the triangulating braces around each
joint. Of course, with the passage of time these braces may have been cut
through during alterations to the building and this may have compromised the
stability of the structure.

Fig. 2.7 Without bracing (a) the simple frame can distort (‘racking’) in
response to horizontal forces. Bracing to create a series of triangles (b)
stiffens the frame.

When iron was introduced it became possible to manufacture complex castings


to act as the joints into which timber members were fixed. Whilst the joints in
such a structure are rigid and the members are fixed at right angles to each
other, the frame can still distort, depending on the relative stiffness of the posts
and beams, as shown in Fig. 4.8.

Unit 2 Heriot-Watt University


Fig. 2.8 With rigid joints the frame deflects in response to a horizontal
force. In (a) the columns are stiffer than the beam, which bends, whilst
in (b) the beam is stiffer than the columns, which bend.

With steel, reinforced concrete and modern timber frames it is possible to


create fixed joints by welding or carefully designed bolting, by ensuring the
continuity of reinforcing steel or by using nailed gusset plates, respectively,
and this creates the rigid knees in a portal frame, referred to by Reid (Appendix
B), p.18. However, recognising that the wind can blow in any direction and
seismic movement in earthquake zones can occur in any direction, some
bracing is probably still necessary to ensure stability, e.g. Reid (Appendix B),
Figs 1.32 and 1.37.

A truss is a framed structure which is designed to span longer distances than a


beam or slab can. Reid (Appendix B) p11-12 introduces the concepts and in
Fig. 2.9 the joints (nodes) are labelled, with the load (UDL) acting vertically
downwards presented as a series of forces above each node, and the reactions
at the supports. Some members are ties which act in tension and these could be
thin (even wires), whilst others are struts which act in compression and need to
resist buckling. We need to identify which is which.

Fig. 2.9 A basic truss.

By analogy with the forces in a beam, the top horizontal members (AC, CE,
EG, etc) are in compression and the bottom members (BD, DF, etc) are in
tension. Identifying which vertical and diagonal members are in tension and
which are in compression can be done by thinking about what happens to the

Heriot-Watt University Unit 2


2: Introduction to Structures and Structural Behaviour

truss if the members are cut through. The sequence in Fig. 2.10 explains this.

Fig. 2.10 Identifying compression and tension members in the truss.

Since the truss is symmetrical, the thought process can stop at this stage and
the complete analysis of members is shown in Fig. 2.11.

Unit 2 Heriot-Watt University


Fig. 2.11 Compression (c) and tension (t) members in the truss.

The size of the forces in each member, and hence the size of the members
themselves needed to resist those forces, can be estimated by resolving them at
each joint using the triangle of forces (or a polygon where more than 3
members meet). In Fig. 2.12 lengths of the arrows show the size of the force in
the members and the angles are the same as those in the truss.

Fig. 2.12 Using the triangle or polygon of forces for the truss. For
equilibrium the triangle/polygon must close. Notice the relatively large
compression force in member GF.

ITQ 2.58.
Referring to Fig. 4.7, is it also necessary to brace the corners in the horizontal
plane?

ITQ 2.59.
In what way are the effects of wind loads and seismic ground forces
analogous?

ITQ 2.60.
Use the process sketched in Fig. 4.10 to identify which members are in tension
and compression in the truss shown below, which is a through-truss typical of
many railway and road bridges, where the traffic load is on the bottom
members as the vehicles pass through the middle.

Heriot-Watt University Unit 2


2: Introduction to Structures and Structural Behaviour

2.7 Beckmann and Bowles – Appendix C


Study Appendix C sections 1.3 to 1.5 and work through the following ITQs.

ITQ 2.61. (p.19)


Dark surfaces absorb solar radiation and their temperature rises. In a famous
modern cathedral the roof consisted of panels fitting between sloping
reinforced concrete beams. The roof panels had a dark aluminium sheet as the
external surface, bonded to a layer of thermal insulation, supported by
reinforced concrete panels fixed between the principal beams. Why did the
aluminium sheet fail and allow water to leak inside?

ITQ 2.62. (p.20)


Why does timber warp when exposed to moist conditions on one side only?

ITQ 2.63. (p.21)


Why is it more practicable to release the dimensional movement by allowing
structural components to expand and contract freely than to restrain it by fixing
them tightly?

ITQ 2.64. (p.22)


Think about how robust the following structures are and how to improve them:
(i) A rectangular steel frame with lightly bolted joints
(ii) A tall, single storey brickwork building such as a sports hall
(iii) A dwelling converted into a shop by creating an open-plan ground floor
(iv) A high level walkway across an atrium space
(v) A high-rise tower comprising reinforced concrete panels.

2.8 Conclusion
This unit has presented the basic principles of structural behaviour in a
qualitative, descriptive way.

2.9 References
Beckmann P, Bowles R (2004) Structural aspects of building conservation,
Elsevier Butterworth-Heinemann.
Reid E (1984) Understanding buildings, Construction Press

Unit 2 Heriot-Watt University

You might also like