You are on page 1of 48

STRENGTH OF MATERIALS

LECTURE 1 – MAIN ENGINEERING MATERIALS

1.1 Concrete and Reinforced Concrete


1.2 Timber
1.3 Steel
1.4 Masonry
1.5 Plastics

1.1 Concrete and Reinforced Concrete

1. INTRODUCTION

Reinforced concrete has been accepted as the most widely used civil engineering material. Concrete in
itself has been proved to be weak in tension but reasonably good in compression. This shortcoming of
concrete as widely-used construction material has been to take care of the tensile handicap of concrete,
and to form in conjunction with each other the composite material called reinforced concrete. The
major constituent of concrete is natural aggregate which is available in different shapes and sizes.
However, artificial aggregates which include blast furnace slag, expanded clay, steel shots, broken
bricks and many other aggregates produced by industrial companies can be also used. The other
constituent of concrete is the material used in putting the aggregates together. This is usually referred
to as the binding medium. For most concretes, the binding medium is formed as a result of the
chemical reaction which takes place between water and cement mixed at an appropriate ratio. Other
more expensive binding media with higher ultimate compressive strengths are also used for specialized
construction such as long-span bridges, nuclear power station, under-ground buildings etc.
Polymer concrete which is better known or sometimes referred to as rein-additive concrete is generally
not suitable for use where the structure is sensitive to fire resistance because it is usually highly
inflammable. However, its resistance to chemical attack makes it the most preferred type of concrete
used in the storage of chemicals and construction of chemical prone drains.
Concrete is also used in conjunction with other materials in a form referred to as composite
construction. The usage of concrete in construction cuts across a wide range of engineering works
which include pavement of airport runways, bridges, roads, dams, chimneys, hydraulic structures,
offshore structures, cooling towers, railway sleepers, ... etc. Concrete is also used structurally in
buildings for columns, foundations, beams, slabs, trusses and precast concrete production. The impact
strength of the normal concrete can be improved by the addition of fibres to the concrete mix. The
commonest fibres used include steel fibres, glass fibres, polypropylene and asbestos. It has been
proved that when fibres of appropriate length to depth ratio are added to a concrete mix the flexural
strength of the resultant concrete is usually improved. Steel fibres reinforced concrete just as polymer
concrete are used mainly for specialized concrete works.

2 CONCRETE

In the past, most of the structures were made either in masonry, steel or timber depending upon the
availability of the materials and the nature of the structure. Recently concrete, as building material, has
come into being. In a very short period, concrete has gained so much importance that today more than
65 per cent of structures coming up in the world are constructed with concrete.
Concrete is composed of organic materials called aggregate such as gravel, sand, crushed stone, slag,
etc, cemented together with some cementing material (most commonly used Portland cement) and
water. When the above materials are mixed together so as to form a workable mixture, it can be given
any shape such as beams, slabs or columns. Cement and water react chemically to bind sand (know as
fine aggregate) and gravel or broken stone (know as coarse aggregate). When the constituent materials
are mixed together, they get hardened after a few hours and a hard stone-like product is obtained. This
is called "concrete". Concrete is very strong in compression but extremely weak in tension and is a
brittle substance. The properties of concrete are dependent upon several factors such as amount of
cement, fine and coarse aggregates, water/cement ratio, temperature at the time of mixing, humidity
during moulding and temperatures and humidity maintained subsequently. The process of keeping the
concrete moist at desired temperature is called "curing of concrete".

3 REINFORCED CONCRETE

Concrete being extremely weak in tension, has the tendency to crack when tensile stresses are induced
in any part of the concrete. To arrest these cracks, reinforcement is provided in the tension zone. It
may, however, be noted that reinforcement does not avoid cracks but it simply arrests further widening
of cracks. So by providing tensile reinforcement at a proper place, the micro cracks do develop but
generally they are so small in width that these are not visible to the naked eye. Such a concrete, which
has reinforcement in tension zone, is known as "Reinforced Concrete". Concrete develops in fact, a
very good bond with steel reinforcement.
The ratio between the amount of fine and coarse aggregate necessarily depends on the grading and
another characteristics of the materials in order that the volume of sand is sufficient to fill the voids in
the coarse aggregate and to produce a dense concrete. Water used for mixing and curing should be
clear and free from injurious amounts of oils, acids, alkalis, salts, sugar, organic materials or other
substances that may be deleterious to concrete and steel reinforcement. Potable water is generally
considered satisfactory for mixing, and curing concrete. Mixing or curing of concrete with (for
example) sea water is not recommended because of the presence of the harmful salts in the sea water.

1.2 Timber

1 INTRODUCTION

By its nature, wood is anisotropic material (i.e. having different properties in different directions).
Generally in materials sciences, relationships between load, duration of load, deformation, and material
strength are usually based on the assumption that the structural material is homogeneous and isotropic,
in both elastic constants and strength properties. Wood is neither, because of its cellular structure and
growth characteristics. The growth of trees is by the addition of cells under the bark (hard substance
that covers a tree), at the ends of branches and at the roots. The thickness and structural characteristics
of the new cell layer depend upon many factors such as temperature, moisture, and species. Most cells
are oriental along the longitudinal axis of the tree while fewer cells develop radially and usually none
tangentially.

2 MAIN PROPERTIES AND SPECIFICATIONS

The growth pattern, in certain climates, leads to the development of annual dark and light rings. These
rings exhibit different strength properties, the summer wood being darker in color and stronger than the
lighter-colored spring wood. In estimating the strength properties of some woods the number of annual
2
rings per inch is used as partial guide. The addition of cells occurs, overall, in a uniform manner and
creates the grain appearance in wood. Wood is often assumed to be an orthotropic material (Fig 5.1)
with the three principal elasticity directions coinciding with the longitudinal (a), radial (r), and
tangential () directions in the tree.

Fig .1 Multilayer tree element with the three principal elasticity directions (a), (r), and 

As with most materials there is inherent variability in the strength of small, clear samples of wood
under short-term loading. Added to this variability are the effects of duration of load and strength-
reducing factors such as knots (knot = a round hard place in a piece of wood where a branch grew). In
addition, wood exhibits directional properties when subjected to various stress states. The strength
properties to consider are associated with normal and shear stresses parallel to the grain, perpendicular
to the grain radially, and perpendicular to the grain tangentially (Fig 1). The difference in strength
properties in the radial and tangential directions is seldom of significance in design. Thus, it is
necessary only to differentiate between directions normal and parallel to the grain. Working stresses
for timber are approximately 20 to 75 percent of the ultimate strength. The large variation of the factor
of safety is due to the variability of the strength properties, which depend on knots, moisture content,
grain, density, voids, splits (cuts or breaks in the material), and many other factors. Wood is not ductile
like steel but, it does have a high capacity of energy absorption which is the consequence of applied
load. The modulus of elasticity of wood varies with species of the same wood, its value being different
in the longitudinal, radial and tangential directions. In concrete design, cracking, creep, dead weight
and ductility are of concern but in the case of timber the design of connections is of utmost importance
and hence the need for an understanding of the strength properties of wood to be used for a particular
construction purpose.
Any timber which is used for structural purposes should be stress graded. Two methods of stress
grading are available. It can be carried out manually by specially trained inspectors. They assess
individual pieces of timber for the number and position of such defects as knots and splits, and then, if
suitable, each piece is stamped with its appropriate grade. There are two manual grades – GS (general
structural) and SS (special structural). Knowing the timber species and the stress grade it is then
possible to place the timber into a strength class, which determines the permissible stresses to be used
in structural design calculations.
The second stress grading method is by machine. Each piece of timber is fed through a machine which
measures the force required to bend the timber. A relationship between the bending stiffness and the
appropriate stress class is then assumed, and the piece stamped accordingly.

3
For softwoods the strength classes range from C14 to C30, where the number refers to the ultimate
bending stress in N/mm2. For example, two of the most commonly used timbers – whitewood and
redwood – are both classed as shown in table 1.

Table 1 Strength classifications for whitewood and redwood timber


Stress grade Strength class
GS C16
SS C24

In terms of durability when selecting a structural timber it is important that it remains intact and fully
functional for a reasonable lifetime. Timber is subjected to decay both fungal attack and wood-boring
insects. (Fungus or funguses = any of those plants without leaves, flowers, or green coloring matter,
growing on other plants or decaying matter, including molds, rusts, mildews = funguses that form a
white coating on things exposed moisture, mushrooms = edible funguses with a stem and domed cap,
noted for its rapid growth), … ect. Wet rot is a fungal attack which occurs when the wood is actually
wet, whereas the more sinister dry rot occurs when the moisture content is above 20%. Dry rot is
particularly troublesome because it can spread rapidly, even through brick-work, by releasing long
tendrils (tendril = a thin curling part of a plant serving to attach itself to a support).

3. CLASSIFICATION

Timber species are usually classified into two main types – hardwoods and softwoods. However, this
classification is independent of the texture of a sawn timber section. Hardwoods originate from broad-
leaved trees such as oak, ask, and mahogany, whereas softwoods originate from coniferous trees such
as spruce, pine, and Douglas fir. Softwoods tend to be quicker growing than hardwoods, and
consequently are cheaper to buy. Most structures are, for this reason, manufactured in softwood. The
selection of timber for usage is usually governed by the code requirements of the British Code for the
Structural Use of Timber in Buildings BS 5268:1996. This is one of the few standards which, is still
written in terms of permissible stresses. However, Eurocode 5 uses a limit state approach. Some
commonly used timber species include Afromosia, a hardwood of West Africa with a density of about
700kg/m3 in seasoned conditions and Agba, another hardwood of West Africa with a density of about
500kg/m3. It has strong resistance to decay and can be used as a structural timber. Other species
include Balsa Beech, Birch, Idigbo, Cedar, Fir, Mahogany, Opepe, Oak, Iroko, Sapele and many other
timber species.

1.3 Steel

1 INTRODUCTION

Steel as other materials is also a widely used engineering material made from iron and it is an alloy
mixed with carbon, chromium, cobalt, copper, manganese, molybdenum, nickel, phosphorus, silicon
and sulphur. Different names have been assigned to steel, based on its composition.
We shall, therefore, in this present Lecture 4 examine the classification and some of the main
properties of steel. Steel is fabricated into a variety of structural shapes for use as beams, columns,
plates, connectors, and to act as reinforcement in the comparatively weak tensile zones of concrete
structures.

4
2. CLASSIFICATION

The commonest classification shows that steels may be grouped into three distinct categories, carbon
steel, alloy steel and structural steel.

2.1 Carbon Steel

Carbon steel is usually classified in three groups:


(1) Low-carbon steel - less than 1/4 percent carbon
(2) Medium carbon steel - 0.25 to 0.5 percent carbon
(3) High carbon steel - over 0.5 percent carbon

2.2 Alloy Steel

Alloy steel is usually made up of a combination of other minerals and metals listed before. For
example should brittle steel be needed, sulphur, silicone and phosphorus may be added proportionately
to the alloy. However, should strength be the optimum requirement, nickel may be added. The
hardness of steel can be sufficiently increased by the addition of manganese. The consequence of this
addition will be the improvement of the steel's abrasion resistance. If copper and chromium are added
to steel, then the steel will have the tendency of resisting corrosion. Another conventional name that is
commonly associated wit alloyed steel is high strength-low alloy steel.

2.3 Structural Steel

Structural steel is another classification that steel is usually associated with. This name is associated
with forms of steel alloy that are used in building structures and bridges. Apart from being strong, it
should have permissible values of modulus of elasticity, compressive and tensile strengths, shear
strength. Generally, structural steel is manufactured using diversified methods; the most commonly
used being the cold-rolled or hot-rolled method. In reinforced concrete structures, reinforcement can be
mild steel bars, ribbed tor steel and steel fibres. Steel is generally used because of its easy availability,
high strength, good bond with concrete and its coefficient of expansion being nearly same as for
concrete. Different types of steel bars are available such as Mild Steel, Hot Rolled Deformed bars,
Cold Twisted bars, Hard drawn steel wire and Rolled Steel.
In pre-stressed concrete structures high strength steel is mostly used as pre-stressing material. High
strength steels, as already said are often obtained by alloying, which permits the manufacture of such
steels under normal operation. Carbon is an economical element for alloying, since it is cheap and easy
to handle. Other alloys include manganese and silicon. The most common method for increasing the
tensile strength for pre-stressing is by cold-drawing, the process of cold-drawing tends to realign the
crystals, and the strength is increased by each drawing so that the smaller the diameter of wires, the
higher their ultimate unit strength. High strength steel for pre-stressing takes one of the three forms:
wires (5-12mm diameter), strands (2-4mm diameter) or bars. For post-tensioning, wires are widely
employed; they are grouped in parallel into cables. Strands are fabricated in the factory by twisting
wires together. High strength steels usually contain 0.7 - 0.8 per cent carbon, 0.6 per cent manganese
and about 0.1 percent silica. Bars are first hot rolled and subsequently heat treated. Cold drawn wires
are tempered to improve their ductility properties. Tempering or stress relieving of wires at 150 -
400C results in increasing of tensile strength. The wires used individually or in wire cables are
generally 5 -7 mm in diameter having ultimate tensile strength of about 1500 N/mm 2. The strands used

5
vary in nominal diameter from 10 to 44 mm. The ultimate tensile strength of high strength steels varies
inversely to diameter of wire.
Structural-steel shapes are manufactured to certain tolerances with respect to dimensional variations
such as camber, cross-section, diameter, square-ness, flatness, length, straightness, sweep, thickness,
weight, and width. The specific limitations are contained in the code requirements of the British Code
for Delivery of Rolled Steel Plates, Shapes, Sheet Piling, and Bars for Structural Use.

1.4 Masonry

1 MAIN DEFINITIONS

Masonry is the art of shaping and laying bricks, burned-clay units, blocks of stones, structural clay
tiles, concrete units, to form walls, columns, slabs, using mortar and grout.
Burned-clay units include common (building brick) and face brick, hollow clay tile, and ceramic tile.
There are many types of finishes to meet beautiful and nice architectural appearance requirements.
Brick masonry units include different grades building bricks made from clay or shale, hollow brick,
sand-lime building brick. Light traffic paving brick, concrete building brick. Hollow and solid load-
bearing concrete masonry units, structural clay load-bearing wall tile, unburned-clay units and glazed
structural clay facing tile are also widely used.
Mortar consists of a mixture of cementitious materials and aggregate to which sufficient water and
approved additives, if any, have been added to achieve as workable, plastic consistency. Bond is more
important to the proper functioning of masonry than the strength of mortar itself is.
Grout consists of a mixture of cementitious materials and aggregate to which water has been added
such that the mixture will flow without segregation of the constituents. Grout is used to fill spaces in
masonry to increase the net area of a section, but more importantly to bond reinforcing into the
masonry.

2 CLASSIFICATION

Masonry is divided into several branches each of which is also subdivided. The main classifications
are:
(i) Constructional masonry
(a) Reinforced Masonry
(b) Unreinforced Masonry
(ii) Monumental masonry
(iii) Marble masonry
(iv) Granite masonry

Each of the fore-listed classification of masonry is linked with a craftsman who has a good mastery of
the aspect he is involved in and is well apt with the techniques involved. The craftsmen include the
mason or stone cutter, machinist, waller, fixer, mould cutter, and many other craftsmen. They are
subsequently divided into groups which include those working on soft stone, marble, Portland stone,
hard-stone, granite, etc

6
1.5 Plastics

1 COMPOSITION

Resin is generally applied to the initial polymeric substance from which all plastic products are made.
Plastic refers to a large group of synthetic materials which are made from a number of common
substances such as coal, salt, oil, natural gas, cotton, wood and water. From these said common
substances, relatively simple chemicals known as monomers, which are capable of reacting with one
another are produced. These said monomers are then built up into chain-like molecules called
polymers.
Resins are obtained naturally and synthetically. Natural resins include resin obtained from trees, amber
- a natural yellow or brownish-yellow resin used for making ornamental objects, and lac- a natural
resin obtained from the exudate of a tree lining insect. Synthetic resins include vinyl, epoxy silicone,
phenol-formaldehyde, lurea formaldehyde, furfural formaldehyde and allyl resin.
Plastic can be made to be thermoplastics or thermosetting plastics. Thermoplastics are those that have
elasticity property. Some thermoplastic products include insulating covering or electrical wire and
sheet goods. Thermosetting plastics are however those that have rigidity property and are permanently
set. For example plastic fittings are produced (made) from thermosetting plastics.

2 PROPERTIES

Generally plastics have extraordinary quality properties that make them unique materials, and these
said properties include abrasion resistance, adhesive qualities, dimensional stability, ductility, high
impact resistance, low moisture absorption, resistance to chemicals, resistance to weathering,
toughness, transparency, lightness quality and foaming quality, resistance to corrosion.

2.1 Polystyrene and PVC

Improvements in chemical technology led to an explosion in new forms of plastics. Among the earliest
examples in the wave of new plastics were "polystyrene" (PS) and "polyvinyl chloride" (PVC),
developed by [IG Farben] of Germany.

Polystyrene is a rigid, brittle, inexpensive plastic that has been used to make plastic model kits. It
would also be the basis for one of the most popular "foamed" plastics, under the name "styrene foam"
or "Styrofoam". Foam plastics can be synthesized in an "open cell" form, in which the foam bubbles
are interconnected, as in an absorbent sponge, and "closed cell", in which all the bubbles are distinct,
like tiny balloons, as in gas-filled foam insulation and floatation devices.

PVC has side chains incorporating chlorine atoms, which form strong bonds. PVC in its normal form is
stiff, strong, heat and weather resistant, and is used for making plumbing, gutters, house siding,
enclosures for computers and other electronics gear. PVC can also be softened with chemical
processing, and in this form it is used for shrink-wrap, food packaging, and raingear.

2.2 Nylon

The real star of the plastics industry was "polyamide" (PA), far better known by its trade name
'''nylon'''. Nylon was the first purely synthetic fiber, introduced by Du Pont Corporation in New York
City. Du Pont's work led to the discovery of synthetic nylon fiber, which was very strong but also very

7
flexible. The first application was for bristles for toothbrushes. However, his real target was silk,
particularly silk stockings.
Nylons still remain important plastics, and not just for use in fabrics. In its bulk form it is very wear
resistant, particularly if oil-impregnated, and so is used to build gears, mechanical bearings, and
because of good heat-resistance, increasingly for under-the-hood applications in cars, and other
mechanical parts.

2.3. Common plastics and their uses

 Polypropylene (PP): Food containers, appliances, car fenders.

 Polystyrene (PS): Packaging foam, food containers, disposable cups, plates, CD and cassette boxes.

 High impact polystyrene (HIPS): fridge liners, food packaging, vending cups.

 Acrylonitrile butadiene styrene (ABS): Electronic equipment cases (e.g., computer monitors, printers,
keyboards).

 Polyethylene terephthalate]] (PET): carbonated drinks bottles, jars, plastic film, microwavable
packaging.

 Polyester (PES): Fibers, textiles.

 Polyamides (PA) Nylons: Fibers, toothbrush bristles, fishing line, under-the-hood car engine
mouldings.

 Polyvinyl chloride (PVC): Plumbing pipes and guttering, shower curtains, window frames, flooring.

 Polyurethanes (PU): cushioning foams, thermal insulation foams, surface coatings, printing rollers.

 Polycarbonate (PC): Compact discs, eyeglasses, riot shields, security windows, traffic lights, lenses.

 Polyvinylidene chloride (PVDC) Saran plastic: Food packaging.

 Polyethylene (PE): Wide range of inexpensive uses including supermarket bags, plastic bottles.

3 MANUFACTURING METHODS

Plastic products are formed by a number of methods which include injection moulding, expandable
bed moulding, compression moulding, transfer moulding, rotational moulding, form moulding, and
blow moulding.
Plastic has, to some extent, added a new dimension to the construction industry and a new dimension
to anticorrosion building materials. Since plastic products are in finished form, it becomes imperative
to develop new skills for their installation. Professional demands that craftsmanship requires has been
sufficiently reduced by the simplicity of plastic installations and their intrinsic (belonging to the
essential nature of plastic products) properties.

8
LECTURE 2 - STRENGTH OF MATERIALS, MAIN BASIC LAWS AND CONCEPTS.

2.1 Linear and Nonlinear Material Behavior


2.2 Newton's Laws
2.3 Types of Forces and Loads
2.4 Types of Supports
2.5 Degrees of Freedom
2.6 Equilibrium Principle
2.7 Section Properties

2.1. Linear and Nonlinear Material Behaviour (Reminder)


The strength of a material, whatever its nature, is defined largely by the internal stresses or intensities
of forces in the material and its load bearing capacity. The bearing capacity could either be in tension
or in compression. One of the simplest loading conditions of a material is that of tension, in which the
fibres of the material are stretched. In the case of metals, the ability of the metal to withstand
deformation without fracture on application of load is of utmost importance. Examination of this
property of metal is usually carried out by subjecting it to tensile strength test, noting in each case the
load that is causing the elongation. A knowledge of these observations lead to the determination of
properties as tensile strength, modulus of elasticity, yield proof stress; percentage elongation of the
material, hardness or resistance to abrasion. Consider, for example, a long steel wire held rigidly at its
upper end, (Fig 2.1) and loaded by a mass hung from the lower end.

Fig. 2.1 - Stretching of a steel wire under end load

If vertical movements of the lower end are observed during loading it will be found that the wire is
stretched by a small, but measurable, amount from its original unloaded length. The material of the
wire is composed of a large number of small crystals which are only visible under microscopic study;
these crystals have irregularly shaped boundaries, and largely random orientations with respect to each
other; as loads are applied to the wire, the crystal structure of the metal is distorted.
For small loads it is found that the extension of the wire is roughly proportional to the applied load, but
for high intensities the extension of the wire is no longer proportional to the applied load. The concepts
of linearity elasticity and plasticity form the basic of the theory of small deformations in stressed
materials and these have been well detailed in the course: ”Properties of Civil Engineering Materials”.

2.2 Newton's Laws

Three Newton's laws of motion form the basis of mechanics and structural analysis.
9
Law 1. "The sum of all external forces acting on a body that is at rest is zero .i. e., ∑F = 0".
This law implies that if the body were not at rest but in motion, then it will continue to be in motion
with constant velocity. It is important that such motion will take place without acceleration. In all
structural mechanics, this law helps to explain the concept of equilibrium. The body is assumed to be
"at rest" with all forces at play in the environment of the body balancing each other. For instance, if we
consider a stool with four legs "at rest" (i.e., in equilibrium) then the weight of the stool which is
transmitted through the legs to the ground is balanced by reactions whose magnitude is the same as the
weight acting directly opposite the weight at the interface of the legs and ground.

Law 2. "If a body in motion possesses acceleration, then the sum of all the external forces on that body
is not zero." The magnitude of the acceleration of the body will be related to the magnitude of the sum
of all the forces while the body moves in the direction of the same resultant force. This law forms the
basics of dynamics.

Law 3. "If one body exerts a force on another body then the latter will also exert a force equal in
magnitude but opposite in direction to the first force."
If multiple forces and bodies are involved, the same principle is used though the results become
slightly complicated. This law together with the first law forms the basic for statics and structural
analysis.

2.3. Types of Forces and Loads (Reminder)

In Engineering Mechanics, forces are exactly, like in the courses of physics, categorized into two: body
force and traction force. The body force, for example is a common occurrence experienced when we
suddenly move up or down in a lift. The traction force is characterized by a point of contact between
two bodies, for example by lifting a baby form a cot we have exerted a traction force. A common unit
is used for both body and tractive forces. If a body has a mass of 1kg, for instance, the force it exerts is
9.81N. When forces acting on a system lie in the same plane, they are called coplanar forces. Coplanar
forces are categorized as concurrent and non-concurrent, depending on whether all the forces pass
through one point or not respectively.
Engineering materials are often subjected to external forces usually called loads which can be
categorized into two: point loads and uniformly distributed loads (Fig.2.2). A point load (called also
concentrated load) is usually represented by an arrow indicating the direction in which the load acts.
Examples point loads are the weight of a person acting on a floor, the legs of a chair or a table
transmitting load to the floor, etc.
A uniformly distributed load comprises a number of concentrated loads closely packed together.
Examples of uniformly distributed loads are wind loads, a shelf of books, etc.

10
Fig. 2.2 Loads and reactions on an architectural structure

In practice, some distributed loads are not always uniform. When non-uniformity of loads
approximates to geometric shapes such as triangular or trapezoidal, they are regarded as such (Fig.3.2).

Fig. 2.3 Non-uniformly distributed loads

2.4. Types of supports (Reminder)

All of architectural engineering structures are supported by four kinds of supports, shown in Fig. 2.4
and namely: roller support; pin support, fixed support, and free support.

11
Fig 2.4 Types of supports

2.5 Degrees of Freedom (Reminder)

Degree of freedom for a structure simply means the direction in which the support allows movement
(Fig.2.4).
The roller support is restrained from moving in the vertical direction but can rotate and also move in
the horizontal direction. Therefore the roller support has two degrees of freedom. Similarly, the hinge
support is restrained from moving in the vertical or horizontal directions but can rotate about its axis.
Thus the hinge support has one degree of freedom, that is the freedom to rotate about its axis.
The rigid or fixed support which is restrained from moving in all directions has zero degree of freedom
and the free (end) support which has freedom to move in all three directions and therefore has three
degrees of freedom.

12
2.6 Equilibrium Principle (Reminder)

According to the laws of statics it is normal to check for the equilibrium of a force system algebraically
by resolving the forces into two orthogonal directions (usually the vertical and horizontal directions)
and the conditions for equilibrium in a two-dimensional system can therefore be summarized by the
following three equations:
The sum of the vertical components of all the forces = 0

∑V = 0

The sum of the horizontal components of all the forces = 0

∑H = 0

The sum of the moments of all of the forces about one point = 0

∑M = 0

2.7 Section Properties

The concept of moment of areas is important when computing stability aspects of sections. Sections of
engineering member often take several shapes due to reasons of efficiency or simply aesthetic. Some
common well known sections in engineering are shown in Fig 2.5.

Fig 2.5 Sections of engineering members

13
2.7.1 First Moment of Area

As in elementary statics the first moment of area of a section about an axis is defined as the product of
the area and the perpendicular distance from the axis to the centroid of the section. Thus the first
moments of area of the section shown in Fig 2.6 are: M xx  Ay; M yy  Ax.

Fig 2.6 Area of a rectangular section

2.7.2 Centroid of Sections

From elementary mechanics, the centre of gravity of any figure is known as that point through which
the section can be suspended and balanced. From the definition of the first moment of area, it is
implicit that the centre of gravity can be defined as:

First moment of area M M yy


y, x   y  xx ; x  .
Area A A

2.7.3 Second moment of area

The second moment of area is more commonly known as the moment of inertia I, and is defined as the
product of the area of the section and the square of the perpendicular distance. For the case shown in
Fig 2.7, I xx  Ay 2 ; I yy  Ax 2 .

Fig. 2.7 Circular section


2.7.4 Parallel Axis Theorem

The parallel -axis theorem states that: ”The moment of inertia of a section about an axis is given by the
sum of the moments of inertia through the centroid of the section and the product of the area and the
square of the perpendicular distance between the two axes” i.e., I axis  I centroid  AD 2
where A is the area of section and D is the perpendicular distance between the two axes.

14
EXAMPLES

Example 2.1

Determine the moments of inertia for the sections shown in Fig. 2.8. and prove
that I axis  I centroid  AD2 ,
where A is the area of the section and D is the perpendicular distance between the two axes

Fig. 2.8 for Example 2.1

Solution:

 
b
1 b 1
(1) I x   x dA   x 2 ldx  l x 3 0  lb 3 .
2
3 3
0
(2) Let O be the centroid of section (area A) whose centroid is at a distance D from the y-axis. A
differential element dA is at a distance d from O. The moment of inertia of dA about the x-axis is:
dIaxis  (D  d) 2 dA . For the entire area I axis   (D  d) 2 dA D 2  dA  2D ddA   d 2 dA.

d
2
Now dA  I centroid, i.e., the moment of inertia about x , the centroid of the section.
Also,  ddA  0 since the differential element is referenced from the centroid of the section and in the
resultant the integral vanishes. Hence I axis  I centroid  D 2  dA .

Example 2.2

Determine the moment of inertia of the rectangle about the centroidal axis X1- X1 Shown in Fig. 2.9.

Fig. 2.9 for Example 2.2


15
Solution:

 
1/ 2
1 1/ 2 1
I axis   y dA as before and 
2
 y3bdy  b y3 1/ 2  b13 .
3 12
1/ 2

Example 2.3

Determine the moment of inertia about the y-axis of the rectangle PQRS shown in Fig. 2.10. Given that
PQ = 6cm; PS = 4cm and OP =5cm.

Fig. 2.10 for Example 2.3

Solution
1
I axis  I centroid  AD2  I yy  AD2  (6) (4) 3  24 (7) 2  1304 cm4
3

Example 2.4

Determine the moment of inertia about the x-axis, Ix of the symmetrical channel shown in Fig 2.11.

Fig. 2.11 for Example 2.4

16
Solution:
It is observed that Ix for rectangles 1 and 2 will be given by (1/3) lb3 but since rectangle 3 is remote
from the x-axis we first use the parallel-axis theorem to determine Ix for rectangle 3 alone. To do this
we need to use I centroid for rectangle 3 (ref.: Example 2.3)
1
(I x ) 3  I centroid  AD2  (3)(2) 3  3(2)(7) 2  296 cm4 .
12
Thus for the whole channel
1 1
I x  (3)(2) 3  (2)(8) 3  296  645.3 cm4 .
3 3

17
LECTURE 3 - ANALYSIS OF STRESS AND STRAIN

3.1 Normal and Shear Stresses


3.2 Normal and Shear Strains
3.3 Mohr’s Circle
3.4 Three Dimensional State of Stress and Strain.

Introduction

Stress and strain are important concepts in the consideration of both strength and rigidity. They are
inevitable and inseparable consequences of the actions of load on a structural material. Stress may be
thought of as the agency which resists loads; strain is the measure of the deformations which occurs
when stress is generated.

3.1 Normal and Shear Stresses

3.1.1 Normal and Shear Stresses in the Cartesian Coordinates System

The stress in a structural element is the internal force divided by the area of the cross section on which
it acts. Stress is therefore internal force per unit area of a cross-section, conversely internal force can be
regarded as the accumulated effect of stress. The strength of a material is measured in terms of the
maximum stress which it can withstand – its failure stress. The strength of a structural element is the
maximum internal force which it can withstand. This depends on both the strength of the constituent
material and the size and shape of its cross section. The ultimate strength of the element is reached
when the stress level exceeds the failure stress of the material. Several different types of stress can
occur in a structural element depending on the direction of the load which is applied in relation to its
principal dimension. If the load is coincident with the principal axis of the element, it causes axial
internal force and produces axial stress called normal or direct stress (Fig. 3.1).

Fig. 3.1 Axial load occurs where the line of action of the
applied load (force) is coincident with the principal axis of the
structural element. This causes axial stress.

The stress that is perpendicular or normal to the cross-section of the element is called normal or direct
stress (designated as ) at that point. Fig. 3.2 shows the normal stress in the Cartesian coordinates
system (x, y, z) acting on an element in three directions.

18
Fig. 3.2 Normal stresses in the Cartesian coordinates system

The axial stress in an element is uniformly distributed across a cross section (Fig. 3.3) and is calculated
from the following equations.
P
 a  ; where  a  axial stress, P  axial force.
A
A  Area of cross  sec tion.

Fig. 3.3 Tensile stress a on the cross-section of an element subjected to axial tension. The
intensity of this is normally assumed to be constant across the cross section.

Axial stress can be tensile or compressive. If the size of cross-section does not vary along the length of
an element, the magnitude of the axial stress is the same at all locations. A load is called a bending-
type load if its direction is perpendicular to the principal axis of the element (Fig. 3.4); this produces
the internal forces of bending moment and shear force which cause a combination of bending stress
and shear stress respectively to act on the cross-sectional planes of the element.

19
Fig. 3.4 Bending-type load occurs where the line of
action of the applied force is perpendicular to the
axis of the element. This causes bending and shear
stresses to occur on the cross-sectional planes.



Where the cross-sectional areas of an element is subjected to stresses which are parallel, rather than
normal to the surface (Fig. 3.5), such stresses are termed as shear stresses (designated by ). The shear
stress components in the Cartesian coordinate system (x, y, z) are:

 xy ,  yx ,  yz ,  zy ,  zx ,  xz

Where the first subscript denotes the direction of the normal to the plane on which the shear stress acts
and the second subscript denotes the direction of the shear stress.

Fig. 3.5 Shear stress components in the Cartesian coordinate system

3.1.2 Two-Dimensional Stress System

A two-dimensional stress system is one in which the stresses at any point in a body act in the same
plane. Consider a two-dimensional stress system subjected to tensile stresses and shear stresses as

20
shown in Fig. 3.6. Such a system is a plane stress system with no normal and shear stresses acting on
the z-direction (with the complementary shear stresses  xz   yz  0 ).

Fig. 3.6 Two-Dimensional stress system

The stresses applied to an element of such a system are in equilibrium in the (x, y) plane and the
element is assumed to be of thickness  z measured in the z direction with sides of lengths  x and  y
in the x and y-directions respectively. Observe that, the shear stress  yx and  xy (top and bottom of the
element) constitute a couple of equal magnitude of  yx  x  y  z and  xy  x  y  z about the center 0.
Therefore, for the element to be in equilibrium:  yx  x  y  z   xy  x  y  z , from which  yx   xy .
Similarly in the (x, z) plane and (y, z) plane  xz   zx and  yz   zy .

Examine a prismatic element of triangular cross-section with its three rectangular faces normal to the
x, y and n directions (Fig. 3.6). The forces acting on the elements are as shown in Fig. 3.7 with t being
its thickness in the t-direction.

Fig. 3.7 Forces acting on the element of the thickness  t

21
For the element’s equilibrium, resolving forces in the n-direction.
 n  t  z   x  y  z cos    xy  y  z sin    yx  x  z cos    y  x  z sin   0
and in t-direction:
 nt  t  z   x  y  z sin    xy  y  z cos    yx  x  z sin    y  x  z cos   0
Considering the geometry of the element,
 x   t sin  and  y   t cos 
and with  z common, the above two expressions reduce to
 n   x cos 2    y sin 2   2  xy sin  cos 

 nt  (  x   y ) sin  cos    xy (cos 2   sin 2  )  because

 nt  t   x  t cos  sin    xy  t cos 2    yx  t sin 2    y  t sin  cos   0

  nt  (  x   y ) sin  cos    xy (cos 2   sin 2  ); sin 2   2 sin  cos  ; cos 2   cos 2   sin 2  ;
x y
  nt   sin 2    xy cos 2 
2

Demonstration:
 nt  t  z   x  y  z sin    xy  y  z cos    yx  x  z sin    y  x  z cos   0
 x   t sin  and  y   t cos 

 nt  t   x  t cos  sin    xy  t cos 2    yx  t sin 2    y  t sin  cos   0

 nt  (  x   y ) sin  cos    xy (cos 2   sin 2 )

sin 2   2 sin  cos  ; cos 2  cos 2   sin 2 


x  y
 nt   sin 2    xy cos 2 
2
Using trigonometric identities sin2 and cos2 we finally obtain the expression for the stress
components as
x  y x  y
n   cos 2   xy sin 2 (3. 1)
2 2
 ( x   y )
and  nt  sin 2    xy cos 2  (3.2)
2
The planes on which the shear stress is zero are known as the principal planes. The normal stress on
the principal plane is termed as the principal stress. The principal stress is obtained by differentiating
Equation (3.1) with respect to angle  and equating to zero to obtain
2  xy
tan 2  p  (3.3)
x  y
where θp is the angle of the normal to the principal plane. The value of stresses on this plane is
obtained by defining the sine and cosine of angle 2p with reference to Equation (3.3) and substituting
in Equation (3.1) to arrive at

22
x  y
1, 2   1 ( x   y ) 2  4  2xy (3.4)
2 2
This equation provides two normal stress components namely 1 the maximum principal stress and 2
the minimum principal stress at the point concerned in 2-dimension (1  2). In a similar manner, the
maximum shear stress is derived by differentiating Equation (3.2) with respect to the angle  and
equating to zero, we have
x  y
tan 2 s   (3.5)
2  xy
Comparing Equation (3.5) with (3.3), we observe that
tan 2 s  tan 2 p  1
This implies that the angle  differs by 90o. Therefore the angle of the normals to the planes differs by
45o, and the planes of maximum and minimum shear stress are at 45o to the principal plane. The
practical implication of this is the development of cracks at window and door openings which always
extend at an angle of 45o. The magnitude of shear stress on plane s is obtained by defining sine and
cosine of angle 2s with reference to Equation (3.5) and substituting in Equation (3.2) to arrive at
1, 2   1 ( x   y ) 2  4 2xy (3.6)
2
where τ1 and 2, are the maximum and minimum shear stresses respectively (τ1  2 ).

The maximum and minimum shear stresses can also be expressed in terms of principal stresses by
combining Equations (3.4) and (3.6):
1   1 ( x   y ); and  2   1 ( x   y ) (3.7)
2 2

3.2 Normal and Shear Strains

The dimensional change which occurs to a specimen of material as a result of the application of load is
expressed in terms of the dimensionless quantity strain. This is defined as the change in a specified
dimension, divided by the original value of that dimension. The precise nature of strain depends on the
type of stress with which it occurs.
Axial normal stress produces axial normal strain, which occurs in a direction parallel to the principal
dimension of the element and is defined as the ratio of the change in length which occurs, to the
original length of the element (Fig. 3.8), in a particular direction (designated as  x ,  y ,  z ; ; normal
strains in x, y, z directions respectively).

L
Fig. 3.8 Axial strain   .
L

23
Shear strain is defined as the change in angle between two planes initially at right angle (designated as
xy for shear strain in the x–y plane). In other words, it is the total angle of distortion. For example,
consider the rectangular element with its edges parallel to x and y axes as in Fig. 3.9, the shear strain of
the element is defined as  xy  1   2 .

Fig. 3.9 Rectangular element and its shear stress

Plane strain is a term used to describe the stain system in which the normal strain in say z–direction
along with the shear strains zx and zy are zero. The shear strain xy is assumed negative when the
angle at the origin of coordinates is increased. For instance in the case of Fig. 3.9, xy is positive as the
angle at the origin of coordinates is reduced. Expressions for normal strain (εn) and shear strain (s)
have been derived in standard texts and will be quoted here only.
On the assumption that strains are small, we have the expression for normal strain as:
x  y x  y  xy
n   cos 2  sin 2 (3.8)
2 2 2
and for shear strain
s x  y  xy
 sin 2  cos 2 (3.9)
2 2 2
where  is the angle measured from the positive x-axis to the plane being considered.

Note that Equations (3.8) and (3.9) are very similar in form to Equations (3.1) and (3.2). Indeed they
are identical if we replace normal stress components by the corresponding normal strain components,

and  xy by s .
2
There appears to be a close resemblance between the transformation equations for stress and strain and

as such the result for stresses can be adapted for strains by merely replacing  n by  n and  n by s .
2
Thus the direction of principal strains is given by
 xy
tan 2 p  (3.10)
x  y
and the principal strain by
x  y
1,2   1 ( x   y ) 2   2xy (3.11)
2 2
Similarly, the direction of planes of shear strain is defined by
x  y
tan 2 s  (3.12)
 xy
with a maximum and minimum values at 45o to the principal planes. The strains are defined by

24
1, 2
  1 ( x ) 2   2xy (3.13)
2 2
or
1, 2
  ( x   y ) 2   2xy (3.14)
2

3.3 Mohr’s Circle

The equations for stress transformation and strain transformation at a point can easily be interpreted
graphically thereby enabling one to have greater insight into the various transformation equations. A
quicker solution can often be obtained. The graphical form of representation is known as Mohr’s
Circle.

3.3.1 Stress Application

A geometrical interpretation of Equations (3.1) and (3.2) leads to a simple method of stress analysis.
We have found already that
x  y x  y
n     cos 2   xy sin 2
2 2
 ( x   y )
n    sin 2    xy cos 2 
2

Take two perpendicular axes ,  (Fig. 3.10); on this coordinate system set off the points having
coordinates ( x ,  xy ) and ( y ,   xy ) corresponding to the known stresses in the x- and y- directions.

Fig. 3.10 Mohr’s circle of Stress

The point P and Q correspond to the stress state ( x ,  xy ), ( y ,   xy ) respectively and are
diametrically opposite. The state of stress (, τ) on a plane inclined at an angle  to 0y is given by the
point R. The line PQ joining these two points is bisected by the 0 axis at the point 0’. With a centre at
0’ construct a circle passing through P and Q. The stresses  and τ on a plane at an angle  to 0y are

25
found by setting off a radius of the circle at an angle 2 to PQ, (Fig. 3.10); the angle2 is measured in a
clockwise direction from 0’P. The coordinates of the point R (, τ) give the direct and shearing stresses
on the plane. We may write Equations (3.1) and (3.2) in the forms:
1 1
  (  x   y )  ( x   y ) cos 2   xy sin 2
2 2
1
   (  x   y ) sin 2   xy cos 2
2
Now, square each equation and add, then we have
2 2
 1  2 1  2
  2 ( x   y )     2 ( x   y )   xy
because:
2 2
1  2 2 2 1  2 2 2
 2 ( x   y ) cos 2    xy sin 2    2 (  x   y ) sin 2   xy cos 2 
2
1  2 2 2 2 2
 2 ( x   y ) (cos 2   sin 2 )   xy (sin 2   cos 2)
2
1 
Thus, all corresponding value of  and τ lie on a circle of radius  ( x   y )   2xy , with its
2 
1
 
centre at the point (  x   y , 0), (Fig. 3.10). This circle defining all possible states of stress is well
2
known as ”Mohr’s Circle of Stress”; the principal stresses are defined by the points A and B at which τ
= 0. The maximum shearing stress (τmax), which is given by the point C, is clearly the radius of the
circle.

Example 3.1

At a point of a material the stresses forming a two-dimensional system are:


2 2 2
 x  50 MN / m ;  y  30 MN / m ;  xy  20 MN / m ;  x ,  y ,  xy having the directions shown in
Fig. 3.11. Draw a Mohr’s Circle of stress and deduce the values of the principal stresses and the
maximum shearing stress in the plane of stresses

Fig. 3.11 Stresses on an inclined plane in a two-dimensional stress system

26
Solution:
On the  - τ diagram, construct a circle with the line joining (50, 20) and (30, – 20) as diameter (Fig.
3.12). The intercepts of the circle on the –axis are 1 = 62.4 MN/ m2 and 2 = 17.6 MN/m2 which are
the principal stresses. The maximum shearing stress is the radius of the circle:
1
 max  (62.4  17.6)  22.4 MN/m 2 . This is the maximum shearing stress in the plane of the applied
2
stresses.

Fig. 3.12 for Example 3.1

Example 3.2

At a point of a material the two-dimensional state of stress is defined by


2 2 2
 x  30 MN / m ;  y   10 MN / m ;  xy  20 MN / m . Find the principal stresses and the maximum
shearing stress.

Solution:
On the  - τ diagram, construct a circle with the line joining (30, 20) and (-10, -20) as diameter (Fig.
3.13). The intercepts of the circle on the -axis are: 1 = 38.3 MN/m2 and 2 = -18.3 MN/m2 which are
the principal stresses. The maximum shearing stress is the radius of the circle: τmax = 28.3 MN/m2.

Fig. 3.13 for Example 3.2

27
3.3.2 Strain Application

Strains in any direction can be represented graphically in a similar way to the stress system. We may
write Equations (3.8) and (3.9) in the forms:
1 1 1
  ( x   y )  ( x   y ) cos 2   xy sin 2
2 2 2
1 1 1
  ( x   y ) sin 2   xy cos 2
2 2 2
Square each equation and then add, we have
1 1 1 1
[  ( x   y )]2  [  ] 2  [ ( x   y )]2  [  xy ] 2
2 2 2 2
1 1 1
Thus all values of ε and  lie on a circle of radius [ ( x   y )]2  [  xy ]2 with its centre at the
2 2 2
1
point [ ( x   y ), 0 ].
2
This circle defining all possible states of strain is usually called Mohr’s circle of strain. For given
values  x ,  y and  xy , it is constructed in the following way: two mutually perpendicular axes, ε and
1 1 1
 are set up, (Fig. 3.14); the points P( x ,  xy ) and Q( y ,   xy ) are located; the line PQ joining
2 2 2
these two points is the diameter of the Mohr’s circle of strain.

Fig. 3.14 Mohr’s circle of strain

1
The diagram is similar to the Mohr’s circle of stress, except that  is plotted along the ordinates and
2
1
not . The values of  and  in an inclined direction are given by the points on the circle at the ends
2
of a diameter making an angle 2 with PQ; the angle 2 is measured clockwise. We note that the
1
maximum and minimum values of ε given by 1 and ε2 in Fig. 3.14, occur when  is zero; 1 and ε2
2
are called principal strains, and occur for directions in which there is no shearing strain.

28
3.4 Three-dimensional State of Stress and Strain

In any two-dimensional stress system we found there were two mutually perpendicular directions in
which only direct stresses 1 and 2 acted; these were called the principal stresses. In any three-
dimensional stress system we can always find three mutually perpendicular directions in which only
direct stresses 1, 2 and 3 in Fig. 3.15 are acting.

Fig. 3.15 Principal stresses in a three-dimensional system

No shearing stresses act on the faces of a rectangular block having its edges parallel to the axes 1, 2
and 3 in Fig. 3.15. These direct stresses are again called principal stresses.
If 1  2  3 then the three-dimensional stress system can be represented in the form of Mohr’s
circles as shown in Fig. 3.16.

Fig. 3.16 Mohr’s circle of stress for a three-dimensional system; circle ’a’ is the Mohr’s circle of
the two-dimensional system 1, 2; circle ’b’ corresponds to 2, 3 and circle ’c’ corresponds to
3, 1. The resultant direct and tangential stress on any plane through the point must correspond
to a point P lying on or between the three circles.

29
Circle ’a’ passes through the points 1 and 2 on the -axis, and defines all states of stress on planes
parallel to the axis 3, (Fig. 3.15), but inclined to axis 1 and axis 2; similarly, circles ’b’ and ’c’ (Fig.
3.16) define stresses on planes parallel to axis 1 and axis 2, respectively. Circle ’c’ having a radius
1
(1   3 ) , includes the two smaller circles. For a plane inclined to all three axes the stresses are
2
defined by a point such as P within the shaded area in Fig. 3.16. The maximum shearing stress is
1
 max  ( 1   3 )
2
and occurs on a plane parallel to the axis 2.

We should note that when one of the principal stresses 3 say, is zero, (Fig. 3.17), we have a two-
dimensional system of stresses 1 , 2; the maximum shearing stresses is the planes 1-2, 2-3, 3-1 are,
respectively,
1 1 1
(1   2 ), 2 , 1.
2 2 2

Fig. 3.17 Two-dimensional stress system as a particular case of a three-dimensional system with
one of the three principal stresses equal to zero.

Suppose, initially, that 1 and 2 are both tensile and that 1  2; then the greatest of the three
maximum shearing stress is
1
1
2
which occurs in 1–3 plane. If on the other hand, 1 is tensile and 2 is compressive, the greatest of the
maximum shearing stresses is
1
( 1   2 )
2
and occurs in the 1–2 plane.

The direct strains corresponding to 1, 2 and 3 for an elastic material are found by taking account of
the Poisson ratio effects in the three directions; the principal strains in the directions 1, 2, 3 are
respectively,

30
1
1  (1    2    3 )
E
1
 2  ( 2    3  1 )
E
1
 3  (  3  1   2 )
E
where  is Young’s modulus, and  is Poisson’s ratio.

The strain energy stored per unit volume of the material is


1 1 1
U  1 1   2  2   3  3
2 2 2

In terms of 1, 2 and 3 this becomes


1
U ( 12   2 2   33  2 1  2  2  2  3  2 3 1 ) .
2E

31
LECTURE 4 - ELASTIC AND PLASTIC ASPECTS OF BEAMS’ BENDING; COMBINED
STRESSES

4.1 General Theory of Elastic Bending


4.2 Distribution of Elastic Bending Stress
4.3 Calculation of Elastic Bending Stress
4.4 General Concepts of Plastic Bending Stress
4.5 Combined Stresses Produced by Axial and Flexural Loads

The application of a transverse load to a structural member invariably produces bending in the
member. The bending process induces stresses and strains in the material from which the beam is
made. The behaviour of beam structures subjected to static transverse loads is to be studied in this
present chapter. Extended attention will be paid to combined stresses produced by axial and flexural
loads.

4.1 General Theory of Elastic Bending

The simplest case to consider is that of a straight beam with a uniform rectangular cross-section
throughout its length. In the initial condition before any load is applied, the side elevation of the beam
viewed along its longitudinal axis could be described as shown in Fig. 4.1.

Fig. 4.1 Side elevation of a beam viewed along its longitudinal axis

In a pre-determined loaded condition where two concentrated couples of equal magnitude but opposite
directions are applied to the ends of the beam, the beam will bend to produce a uniform radius of
curvature R when opened in the y–z plane. The side view of the beam would then become that shown
in Fig. 4.2.

Fig. 4.2 Side view of a bent beam with a radius of curvature R when opened in the y–z plane

A cross-sectional view of the rectangular cross-section of the beam can be drawn in fig. 4.3 which
shows the position of the centroid C of the cross-section, with the horizontal x-axis and vertical y-axis
both passing through the centroid. It is evident that C-x and C-y are both axes of symmetry of the
cross-section.

32
Fig.4.3 Cross-sectional view of the rectangular cross-section of a beam

Fig. 4.2 shows that in general, the long dimension or length of the beam is bent in y–z plane, although,
there is one part of the beam which is unaffected by the curvature of beam. This is the longitudinal
centroidal axis of the beam, and the length of this axis is unchanged in length. In the case illustrated
above, the top surface of the beam is stretched, while the lower surface is compressed by the bending
action of the beam.
A small element of the beam which is remote from the ends will now be considered. Sections AB and
FD (Fig. 4.1) are the ends of the small element, and they are parallel to each other when the beam is in
an unloaded condition. When the load is applied and the beam is viewed in side elevation, it is
assumed that plane sections remain plane in the bent beam.
Section A’B’ and F’D’ (Fig. 4.2) are sections through the bent beam, but they are no longer parallel. In
the bending process, some of the longitudinal fibres, eg. A’F’, are stretched, while others, such as B’D’
are compressed.
The unstrained middle surface is known as the neutral surface, and the line parallel to C-x is known as
the neutral axis (Fig. 4.3).
Strains due to bending increase with increasing distance from the neutral axis.

An elemental fibre HJ (Fig. 4.4) of the beam shown in Fig. 4.1 will now be considered. The fibre HJ is
parallel to the longitudinal C-z axis, and is at a distance y from the neutral surface on the tension side
of the beam.

Fig. 4.4 Longitudinal section of a small element of the beam showing fiber HJ

33
If the length of the fibre HJ in the unloaded condition = z
z
The length of fibre H’J’ (Fig. 4.5) in the loaded condition = (R  y)
R

Fig. 4.5 Longitudinal section of a small element of the beam showing fiber H’J’

Therefore, during bending the fibre HJ extends by an amount.


z y
( R  y)  z  z
R R
Thus, the longitudinal strain ε, in HJ is
y
( z)
y
 R  (4.1)
z R

It can therefore be deduced that the longitudinal strain in any fibre along the length of the beam is
proportional to the value of y, its distance from the neutral surface. In compression fibres (in the lower
half of the beam in this case) the longitudinal strains are negative. If the material of the beam remains
linear-elastic during bending (i.e. the relationship between stress and strain remains linear) then the
longitudinal stress on fibre HJ is:
y
  EE (4.2)
R
where E = Young’s modulus of the material.

4.2 Distribution of Elastic Bending Stress

The distribution of bending stress over the cross-section of the beam can now be determined by
studying the relationship between stress  and distance from the centroidal axis. As shown in Fig. 4.6,
this relationship is linear and implies that the maximum values of bending stress occur on the
outermost fibres of the beam, i.e. those parts of the beam which are most remote from the neutral
(centroidal) surface. This observation is true for all cross-sections subjected to pure elastic bending,
and can be applied with careful interpretation if the cross-section is not symmetrical, or if a
combination of bending and axial force exists within the beam.

34
Fig. 4.6 Distribution of elastic bending stress over the cross-section of the beam

If the equilibrium in the axial direction of a longitudinal section of the beam is considered, the bending
stresses are distributed symmetrically about the C-x axis and therefore there can be no resultant,
longitudinal thrust on the element of the beam.
The stresses on a small part of the cross-section of the beam will now be considered. A diagram of the
rectangular cross-section is illustrated in Fig. 4.7.

Fig. 4.7 Diagram of the rectangular cross-section of the beam

The longitudinal force on the thin strip of the beam section (b wide and y deep) is given by
F   by (4.3)
and the moment produced by this force F about the centroidal axis is
M strip  F y  ( b  y ) y (4.4)
The resultant hogging moment, M, can thus be shown to be
h

2
M   b yy . (4.5)
h

2

Substituting for the expression for  into the Equation (4.5), we can obtain
35
h

2
E E
M 
R h
b y 2  y or M  I xx
R
(4.6)

2
where Ixx is the well known second moment of area of the cross-section about the C-x axis.

Combining Equation (4.6) with Equations (4.5) and (4.2), the previous expressions relating bending
stress and bending moment we can obtain
 E M
  (4.7)
y R I xx
This Equation (4.7) is known as the Fundamental Equation of Elastic Bending and is very important in
the analysis of bending stress.

4.3 Calculation of Elastic Bending Stress

From previous concepts bending moment occurs in an element if the external loads cause bending
moment to act on its cross-sections. The magnitude of the bending stress varies within each cross-
section from maximum stresses in tension and compression in the extreme fibres on opposite sides of
the cross-section, to a minimum stress in the centre (at the centroid) where the stress changes from
compression to tension (Fig. 4.8). It may also vary between cross-sections due to variation in the
bending moment along the element.

Fig. 4.8 Distribution of elastic bending stress on a beam cross-section ABCD

36
(Fig. 4.8) illustrates the distribution of bending stress on a cross-section of a beam element carrying a
bending-type load:
(a) The deflected shape of the compressive stress occurs on the inside of the curve (upper half of the
cross-section) and the tensile stress on the outside of the curve;
(b) The cut-away diagram. Shear force and shear stress are not shown.

The magnitude of bending stress at any point in an element depends on four factors namely:
(i) the bending moment at the cross-section in which the point is situated;
(ii) the size of the cross-section
(iii) the shape of the cross-section
(iv) the location of the point within the cross-section
The relationship between these four factors is:
My
 (4.8)
I
Where  is the bending stress at a distance y from the neutral axis of the cross-section;
M is the bending moment at the cross-section;
I is the second moment of area of the cross-section about the axis through its centroid.

We may note that equation (4.8) is deduced from equation (4.7), the Fundamental Equation of Elastic
Bending.

Now consider the cross-section of a beam shown in Fig. 4.9, a beam subjected to a bending moment,
M, which causes a maximum stress, max, which is less than the yield stress, y.

Fig. 4.9: A rectangular beam subjected to elastic bending

We must satisfy horizontal equilibrium and hence the total tensile force = total compressive force. For
symmetrical sections, therefore, the neutral axis must lie at mid-height. We need to find the
relationship between the stress max, the bending moment, M, and the beam cross-sectional
dimensions, B and D:
pull force = push force = mean stress  area
= max/2  B  D/2
= max  BD/4
For a triangle the centroid is one third of the height from the base, therefore the distance between the
centroids of the two forces:

37
lever arm = 2/3  D

therefore M = max  BD/4  2D/3


= max  BD2/6

or M = max  Z (4.9)

where Z is the elastic section modulus, and equals BD2/6 for a rectangular section. (Take care not to
confuse ’the elastic section modulus’ with ’the modulus of elasticity’. The first is a geometric property,
based on the shape and dimensions of the section. The second is a material stiffness property.

EXAMPLE 4.1

A T-section of uniform thickness 1 cm shown in Fig. 4.10 has a flange breadth of 10 cm and overall
depth of 10 cm. Calculate the allowable bending moments about the principal axes if the bending
stresses are limited to 150 MN/m2.

Fig. 4.10 for Example 4.1

Solution:
Suppose y is the distance of the principal axis C-x from the remote edge of the flange. The total area of
the section is
A = (0.10)(0.01) + (0.09)(0.01) = 1.90  10-3 m2.
On taking first moment of area about the upper edge of the flange,
Ay  (0.10)(0.01)(0.005)  (0.09)(0.01)(0.055)  0.0545  10 3 m 3
Then
0.0545  10 3
y  0.0287 m
1.90  10  3
The second moment of area of the flange about the axis C-x is

38
1
(0.10)(0.01) 3  (0.10)(0.01)(0.0237) 2  0.570  10  6 m 4
12
The second moment of area of the web about the axis C-x is
1
(0.01)(0.09) 3  (0.09)(0.01)(0.0263) 2  1.230  10  6 m 4
12
Then
I x  (0.570  1.230) 10 6  1.8  10 6 m 4
For bending about the axis C-x, the greatest bending stress occurs at the toe of the web, as shown in the
Fig.4.10. The maximum allowable moment is
(150  10 6 )(1.8  10 6 )
Mx   3790 Nm
0.0713
The bending stress in the extreme fibres of the flange is only 60.4 MN/m2at this bending moment. The
second moment of area about the axis C-y is
1 1
I y  (0.01)(0.10) 3  (0.09)(0.01) 3  0.841  10  6 m 4
12 12
The T-section is symmetrical about the axis C-y and for bending about this axis equal tensile and
compressive stresses are induced in the extreme fibres of the flange; the greatest allowable moment is
(150  10 6 )(0.841  10 6 )
My   2520 Nm
0.05

4.4 General Concepts of Plastic Bending Stress

Fig. 4.11 illustrates the idealized stress-strain relationship for mild steel.

Fig. 4.11: Stress-strain relationship for mild steel

In the previous discussion of the elastic bending of beams, the most heavily stressed fibres were those
which were most remote from the centroidal axis of the cross-section.
If the load had been increased sufficiently it can be deduced that the stress on the outer fibres would
reach the value of the yield stress of the material.
Regions of the cross-section which have stress levels at the yield stress of the material are considered
to have become plastic (Fig. 4.12).

39
Fig. 4.12: Stress changes from purely elastic bending to fully plastic bending

As the load increases beyond the point at which yielding first occurs, so the area of plasticity advances
towards the neutral axis of the beam (Fig. 4.12).
If all the cross-section becomes plastic then a plastic hinge is said to have formed, which in many
cases, leads to structural collapse of the beam. Collapse of the beam may not occur if the beam is
statically indeterminate; these special cases will be examined later.
If only part of the cross-section is plastic (Fig. 4.12) then the resistance of the beam can be regarded as
a combination of elastic and plastic bending.

Calculation of Plastic Moment of Resistance of a Rectangular Cross-section

As shown in Fig. 4.12 the rectangular cross-section of breadth b and height h will be considered. It is
assumed that the applied loading condition is such that the top half is completely at yield stress in
tension and the lower half is completely at yield stress in compression, ie. the beam is fully subjected
to plastic bending (Fig. 4.13).

Fig. 4.13: Fully plastic bending stress block

The force acting on the top half of the cross-section of the beam, FP, is
1
FP  (  b  h )   Y (4.10)
2

40
where Y is the yield stress of the material. In Fig. 4.13 this force acts to the right. A force of equal
magnitude but oriented in opposite direction (ie. to the left) acts on the lower half of the cross-section.
Note that these two forces must be of the same magnitude for equilibrium. These two forces form a
couple whose moment is called the plastic moment of resistance. Thus
1 h 1
M P XX   b  h   Y   b h 2  Y (4.11)
2 2 4
From the discussion of elastic bending, the elastic moment of resistance, ME XX of the rectangular
section shown in Fig. 4.9 is that moment which is M, and causes yield stress to be produced only at the
most extreme fibres. Taking into consideration that in Fig. 4.9 the height of the rectangular cross-
section D = h it can be demonstrated from Equation (4.9) that
1
M E XX  b h 2  Y (4.12)
6
Therefore, from the combination of Equations (4.9), (4.11) and (4.12), for a rectangular cross-section,
3
M P XX   M E XX (4.13)
2
Equation (4.13) shows the direct relationship between elastic and plastic moments of resistance for a
rectangular cross-section, and is important in the concepts of structural collapse.

4.5 Combined Stresses Produced by Axial and Flexural Loads

A column-base interface representing an example of combined axial-flexural load action is illustrated


in Fig. 4. 14. The column transmits an axial load P and a bending moment M to the foundation base.

Fig. 4.14: Column-base interface representing an example of combined axial-flexural load action

41
P My
The combined stresses produced by P, and by M, respectively act in the same line and their
A I
effects can be easily summed algebraically.
P My
  (4.14)
A I
Algebraic signs in Equation (4.14) are important. Tensile and compressive stresses are respectively
regarded as positive and negative. The bending moment introduced in Equation (4.7) will take care of
My
the sign of the term ( ).
I

EXAMPLE 4.2
The structural beam PQ shown in Fig. 4.15 is subjected to an inclined point load at the mid-span of the
beam. Calculate the magnitude of stress at top and bottom of the beam at a point R located at a
distance of 3 m from P.

Fig. 4.15 for Example 4.2


Solution:

The inclined 70 kN load is resolved into two components including:


(i) horizontally H = 70 cos 60 = 35 kN
(ii) vertically V = 70 cos 30 = 60.6 kN
The bending moment about point R is MR = 30.3 × 3 = 90.9 kNm

The horizontal component H = 35 kN is shown in Fig. 4.16 and is compressive.

Fig. 4.16: Horizontal component H

42
Thus it is evident that the beam is subjected to a combined compressive axial thrust of 35 kN and a
varying flexural moment which has a magnitude of 90.9 kNm at R.

At the bottom of the beam, the extreme fibres are stretched from the effect of the load component H
and the same fibres are also stretched from the vertical flexural load V. Thus from Equation (4. 14)

P My
 bottom   
A I
35  10 3 90.9  10 6  300
   0.194  5.05  5.24 N/mm 2
300  600 1
 300  600 3
2
 top  0.194 - 5.05  - 4.86 N/mm 2
Thus the distribution of the combined stresses at top and bottom of the beam is shown in Fig. 4. 17

Fig. 4.17: Distribution of combined stresses at top and bottom of the beam

43
LECTURE 5 - COMPOSITE BEAMS

5.1 Engineering Materials for Composite Beams


5.2 Steel Reinforced Timber Beam

5.1 Engineering Materials for Composite Beams

In engineering construction, beams are fabricated from comparatively cheaper materials of low
strength which are reinforced by small amounts of high strength materials such as steel. In this way,
for example a timber of rectangular section may have steel plates bolted to its side or to its top and
bottom surfaces. Also concrete beams are reinforced in their weak tension zones and also, if necessary,
in their compression zones (when the intensity of load is high) by steel reinforcing bars. The design of
reinforced concrete beam and steel-concrete composite beams is covered by Codes of Practice and
relies, as in the case of steel beams on limit state analysis.
The design of steel reinforced timer beams is not covered by a Code, we shall therefore limit the
analysis of this type of composite beam to an elastic approach, using Equation (4.7) which is the
fundamental equation of elastic bending.

5.2 Steel Reinforced Timber Beam

The timber joist section of breadth, b, and depth, d, shown if Fig. 5.1 is reinforced by two steel plates
bolted to its sides, each plate being of thickness, t and depth, d. Let us suppose that the beam is bent to
a radius R, at this section by a positive bending moment M. Clearly, since the steel plates are firmly
attached to the sides of the timber joist, both are bent to the same radius R. Thus, from Equation (4.7),
the bending moment, Mt, carried by the timber joist is
E I
Mt  t t (5.1)
R

Fig. 5.1 Timber joist section reinforced by two steel plates bolted to the sides

In equation (5.1) Et is Young’s modulus for the timber and It is the second moment of area of the
timber section about the centroidal axis Gx. Similarly for steel plates
E I
Ms  s s (5.2)
R
in which Is is the combined second moment of area about axis Gx of the two plates. The total bending
moment is then

44
1
M  M t  Ms  (E t I t  E s I s )
R
from which
1 M
 . (5.3)
R E t I t  E s Is

From a comparison of Equations (5.3) and (4.7) we see that the composite beam behaves as a
homogeneous beam of bending stiffness EI where
EI  E t I t  E s I s
 E  (5.4)
or EI  E t  I t  s I s 
 Et 

The composite beam may therefore be treated wholly as a timber beam having a total second moment
of area
E I
It  s s
Et
This is equivalent to replacing the steel reinforcing plates by timber ‘plates’ each having a thickness,
(Es /Et) t as shown in Fig. 5.2 (a).

(a) (b)

Fig. 5.2 Equivalent timber reinforcing ‘plates’ and equivalent steel ‘joist’

Alternatively, the beam may be transformed into a wholly steel beam by writing equation (5.4) as
E 
E I  E s  t I t  I s 
 Es 
so that the second moment of area of the equivalent steel beam is
Et
I t  Is
Es

45
which is equivalent to replacing the timber joist by a steel ’joist’ of breadth (Et/Es) b (Fig. 5.2 (b)). We
should note that the transformed sections of Fig. 5.2 apply only to the case of bending about the
horizontal axis Gx. We should note also that the depth, d, of the composite beam is unchanged by
either transformation.
The direct stress due to bending in the timber joist is obtained using Equation (4.7), i.e.
Mt y
t  (5.5)
It
From Equations (5.1) and (5.3)
Et It
Mt  M
E t I t  E s Is
or (5.6)
M
Mt 
E I
1 s s
Et It
Substituting in Equation (5.5) from Equation (5.6) we have
My
t  (5.7)
Es
It  Is
Et
Equation (5.7) could in fact have been deduced from Equation (4.7) since It + (Es/Et) Is is the second
moment of area of the equivalent timber beam of Fig. 5.2(a). Similarly, by considering the equivalent
steel beam of fig. 5.2(b), we obtain the direct stress distribution in the steel, i.e.
My
s  (5.8)
Et
Is  It
Es
Example 5.1

A composite beam section is fabricated from two timber joists each 100 mm x 400 mm with a steel
plate 12 mm x 300 mm placed symmetrically between them (Fig. 5.3). If the beam is subjected to a
bending moment of 50 kN.m, determine the maximum stresses in the steel and in the timber. The ratio
of Young’s modulus for steel to that of timber is 12:1.

Fig. 5.3 for Example 5.1


Solution

The second moments of area of the timber and steel about the centroidal axis Gx are

46
400 3
I t  2 100   1067 x 10 6 mm 4
12
3
300
and I s  12   27 10 6 mm 4 respectively.
12

Therefore, from equation (5.7) we have


50 10 6  200
t     7.2 N / mm 2
6 6
1067 10  12  27 10
and from equation (5.8)
50 10 6  150
s     64.7 N / mm 2 .
6 6
27 10  1067 x 10 / 12
Consider now the steel reinforced timber beam of Fig. 5.4 (a) in which the steel plates are attached to
the top and bottom surfaces of the timber. The section may be transformed into an equivalent timber
beam (Fig. 5.4 (b)) or equivalent steel beam (Fig. 5.4 (c) by the methods used for the composite beam
of Fig. 5.1. The direct stress distributions are then obtained from equations (5.7) and (5.8). There is
however, one important difference between the composite beam of Fig. 5.1 and that of Fig. 5.4 (a).

Fig. 5.4 Timber joist section with steel plates attached to the top and bottom surfaces

In the later case, when the beam is subjected to shear loads, the connection between the timber and
steel must resist horizontal complementary shear stresses as shown in Fig. 5.5.

Fig. 5.5 Connection between timber and steel resisting horizontal complementary shear stresses
Generally, it is sufficiently accurate to assume that the timber joist resists all the vertical shear and then
calculate an average value of shear stress, τav. Thus

47
Sy
 av 
bd
so that, based on this approximation, the horizontal complementary shear stress is S y/bd and the shear
force per unit length resisted by the timber/steel connection is Sy/d.

Example 5.2

A timber joist section 100mm  200 mm is reinforced on its top and bottom surfaces by steel plates 15
mm thick  100mm wide. The composite beam is simply supported over a span of 4 m and carries a
uniformly distributed load of 10 kN/m. Calculate the maximum direct stress in the timber and in the
steel and also the shear force per unit length transmitted by the timber/steel connection. Take E s/Et =
15.

Solution

The second moments of area of the timber and steel about a horizontal axis through the centroid of the
beam section are
100  200 3
It   66.7 10 6 mm 4
12
and Is = 2  15  100  107.52 = 34.7 x 106 mm4 respectively. The maximum bending moment in the
beam occurs at mid-span and is
10  4 2
M max   20 kN.m
8

From equation (5.7)


20 10 6  100
 t , max     3.4 N / mm 2
6 6
66.7 10  15  34.7 10
and from Equation (5.8)
20 10 6 115
 s, max .     58.8 N / mm 2 .
6 6
34.7 10  66.7 10 / 15

The maximum shear force in the beam occurs at the supports and is equal to (10  4)/2 = 20 kN. The
average shear stress in the timber joist is then
20 10 3
 av   1 N / mm 2 .
100  200
It follows the shear force per unit length in the timber/steel connection is 1  100 = 100 N/mm or 100
kN/m. We should note that this value is an approximation for design purposes.

48

You might also like