0% found this document useful (0 votes)
128 views92 pages

RC 1 2015 16 Lecture Note

This document provides an overview of reinforced concrete design. It introduces the key differences between plain and reinforced concrete, the advantages and disadvantages of reinforced concrete, and outlines the typical design process. The document discusses different design philosophies and codes, including the limit states method. It also covers materials used in reinforced concrete like concrete and reinforcing steel. Finally, it previews the content to be covered in subsequent chapters, which will focus on limit state design for flexure, shear, and serviceability considerations.

Uploaded by

tekalign
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
128 views92 pages

RC 1 2015 16 Lecture Note

This document provides an overview of reinforced concrete design. It introduces the key differences between plain and reinforced concrete, the advantages and disadvantages of reinforced concrete, and outlines the typical design process. The document discusses different design philosophies and codes, including the limit states method. It also covers materials used in reinforced concrete like concrete and reinforcing steel. Finally, it previews the content to be covered in subsequent chapters, which will focus on limit state design for flexure, shear, and serviceability considerations.

Uploaded by

tekalign
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 92

AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

TABLE OF CONTENTS

CHAPTER 1. INTRODUCTON TO REINFORCED CONCRETE .............................................. 1


1.1. Introduction .......................................................................................................................... 1
1.2. Plain and Reinforced Concrete ............................................................................................ 1
1.2.1. Plain Concrete ............................................................................................................... 1
1.2.2. Reinforced Concrete ..................................................................................................... 2
1.3. ADVANTAGES AND DISADVANTAGES OF reinforced concrete for a structure ....... 4
1.4. The Design Process .............................................................................................................. 5
1.4.1. Objectives of design ...................................................................................................... 5
1.4.2. The design process ........................................................................................................ 5
1.5. Design codes and handbooks ............................................................................................... 5
1.5.1. Purpose of Codes .......................................................................................................... 5
1.5.2. Introduction to Eurocodes ............................................................................................. 6
1.6. Design philosophies ............................................................................................................. 9
1.6.1. Introduction ................................................................................................................... 9
1.6.2. Working stress method (WSM) .................................................................................. 10
1.6.3. Ultimate load method (ULM) ..................................................................................... 11
1.6.4. Limit states method (LSM) ......................................................................................... 12
1.7. Materials ............................................................................................................................ 15
1.7.1. Behavior of concrete under compression .................................................................... 15
1.7.2. Behavior of concrete under tension ............................................................................ 20
1.7.3. Reinforcing steel ......................................................................................................... 22
1.8. EUROCODE’S recommendations for limit states design ................................................. 23
1.8.1. Actions ........................................................................................................................ 24
1.8.2. Material ....................................................................................................................... 30
CHAPTER 2. LIMIT STATE DESIGN FOR FLEXURE ........................................................... 36
2.1. Introduction ........................................................................................................................ 36
2.1.1. Analysis vs. design...................................................................................................... 36
2.1.2. Statics of Beam action ................................................................................................ 36

i
AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

2.2. Distribution of strains and stresses across a section in bending ........................................ 39


2.3. Ultimate limit state for flexure ........................................................................................... 40
2.3.1. Basic assumptions for flexure at the ULS................................................................... 40
2.3.2. Possible range of strain distributions at ULS .............................................................. 40
2.3.3. Limiting Compressive strains at ULS ......................................................................... 41
2.4. Types of flexural failures ................................................................................................... 42
2.5. Analysis of beams for Flexure at the ULS ......................................................................... 43
2.5.1. Analysis of singly reinforced beam sections .............................................................. 44
2.5.2. Analysis of doubly reinforced beam sections ............................................................. 45
2.5.3. Analysis of flanged sections ....................................................................................... 46
2.6. Design of beams for flexure at ULS .................................................................................. 51
2.6.1. Concrete cover ............................................................................................................ 51
2.6.2. Minimum and Maximum area of reinforcement ......................................................... 56
2.6.3. Design of singly reinforced beam sections ................................................................. 57
2.6.4. Design of doubly reinforced sections ......................................................................... 59
2.6.5. Design of flanged beams ............................................................................................. 59
CHAPTER 3. LIMIT STATE DESIGN FOR SHEAR ................................................................ 61
3.1. Theoretical Background ..................................................................................................... 61
3.1.1. Diagonal tension in homogeneous elastic beams........................................................ 61
3.1.2. Behavior of beams failing in shear ............................................................................. 63
3.1.3. Factors affecting the shear strength of beams without web reinforcement ................ 67
3.2. Design of beams for vertical shear according to EN 1992-1-1-2004 ................................ 68
3.2.1. Members not requiring design shear reinforcement ................................................... 69
3.2.2. Members requiring design shear reinforcement ......................................................... 70
3.2.3. Additional tensile force in longitudinal reinforcement ............................................... 71
3.2.4. Minimum area and maximum spacing of shear reinforcement .................................. 71
3.2.5. Procedure for design ................................................................................................... 72
CHAPTER 4. SERVICEBILITY LIMIT STATE ........................................................................ 73
4.1. INTRODUCTION ............................................................................................................. 73
4.2. ELASTIC ANALYSIS OF BEAM SECTIONS................................................................ 74
4.2.1. SECTION UN-CRACKED ........................................................................................ 74

ii
AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

4.2.2. SECTION CRACKED................................................................................................ 74


4.3. STRESS CONTROL ......................................................................................................... 75
4.4. SERVICEBILITY LIMIT STATE OF CRACKING ........................................................ 76
4.4.1. GENERAL INTRODUCTION ................................................................................... 76
4.4.2. CAUSES OF CRACK ................................................................................................ 76
4.4.3. CRACK CONTROL ................................................................................................... 78
4.4.4. CRACK WIDTH CALCULATION ........................................................................... 82
4.5. SERVICEBILITY LIMIT STATE OF DEFLECTION .................................................... 84
4.5.1. GENERAL INTRODUCTION ................................................................................... 84
4.5.2. CASES WHERE CACLCULATION MAY BE OMITTED ..................................... 85
4.5.3. DEFLECTION LIMITS.............................................................................................. 87
4.5.4. CALCULATION FOR DEFLECTION ...................................................................... 87

iii
AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

CHAPTER 1. INTRODUCTON TO REINFORCED CONCRETE

1.1. INTRODUCTION
Traditionally, the study of reinforced concrete design begins directly with a chapter on materials,
followed by chapters dealing with design. In this material, a departure is made from that
convention. It is desirable for the student to have first an overview of the world of reinforced
concrete structures, before plunging into the finer details of the subject. Accordingly, this section
gives a general introduction to reinforced concrete and its applications. It also explains the role
of structural design in reinforced concrete construction, and outlines the various structural
systems that are commonly adopted in buildings.

That concrete is a common structural material is, no doubt, well known. But, how common it is,
and how much a part of our daily lives it plays, is perhaps not well known — or rather, not often
realized. Structural concrete is used extensively in the construction of various kinds of buildings,
stadia, auditoria, pavements, bridges, piers, breakwaters, berthing structures, dams, waterways,
pipes, water tanks, swimming pools, cooling towers, bunkers and silos, chimneys,
communication towers, tunnels, etc. It is the most commonly used construction material,
consumed at a rate of approximately one ton for every living human being. “Man consumes no
material except water in such tremendous quantities”.

1.2. PLAIN AND REINFORCED CONCRETE

1.2.1. PLAIN CONCRETE


Concrete may be defined as any solid mass made by the use of a cementing medium; the
ingredients generally comprise sand, gravel, cement and water. That the mixing together of such
disparate and discrete materials can result in a solid mass (of any desired shape), with well-
defined properties, is a wonder in itself. Concrete has been in use as a building material for more
than a hundred and fifty years. Its success and popularity may be largely attributed to (1)
durability under hostile environments (including resistance to water), (2) ease with which it can
be cast into a variety of shapes and sizes, and (3) its relative economy and easy availability. The
main strength of concrete lies in its compression-bearing ability, which surpasses that of
traditional materials like brick and stone masonry. Advances in concrete technology, during the

CHAPTER 1 - Introduction Page 1


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

past four decades in particular, have now made it possible to produce a wide range of concrete
grades, varying in mass density (1200−2500 kg/m3) and compressive strength (10 −100 MPa).

Concrete may be remarkably strong in compression, but it is equally remarkably weak in tension
[Figure 1-1(a)]. Its tensile strength is approximately one-tenth of its compressive strength.
Hence, the use of plain concrete as a structural material is limited to situations where significant
tensile stresses and strains do not develop, as in hollow (or solid) block wall construction, small
pedestals and ‘mass concrete’ applications (in dams, etc.).

1.2.2. REINFORCED CONCRETE


Concrete would not have gained its present status as a principal building material, but for the
invention of reinforced concrete, which is concrete with steel bars embedded in it. The idea of
reinforcing concrete with steel has resulted in a new composite material, having the potential of
resisting significant tensile stresses, which was hitherto impossible. Thus, the construction of
load-bearing flexural members, such as beams and slabs, became viable with this new material.
Its utility and versatility are achieved by combining the best features of concrete and steel.
Consider some of the widely differing properties of these two materials that are listed below in
Table 1-1.

Table 1-1- Complementary properties of Concrete and Steel

Concrete Steel
Strength in Tension Poor Good
Strength in Compression Good Good, but slender bars will buckle
Strength in Shear Fair Good
Durability Good Corrodes if unprotected
Fire resistance Good Poor, suffers rapid loss of strength at high temperature
It can be seen from this list that the materials are more or less compatible. The steel bars
(embedded in the tension zone of the concrete) compensate for the concrete’s incapacity for
tensile resistance, effectively taking up all the tension, without separating from the concrete
[Figure 1-1(b)]. The bond between steel and the surrounding concrete ensures strain
compatibility, i.e., the strain at any point in the steel is equal to that in the adjoining concrete.
Moreover, the reinforcing steel imparts ductility to a material that is otherwise brittle. In

CHAPTER 1 - Introduction Page 2


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

practical terms, this implies that if a properly reinforced beam were to fail in tension, then such a
failure would, fortunately, be preceded by large deflections caused by the yielding of steel,
thereby giving ample warning of the impending collapse [Figure 1-1(c)].

Tensile stresses occur either directly, as in direct tension or flexural tension, or indirectly, as in
shear, which causes tension along diagonal planes (‘diagonal tension’). Temperature and
shrinkage effects may also induce tensile stresses. In all such cases, reinforcing steel is essential,
and should be appropriately located, in a direction that cuts across the principal tensile planes
(i.e., across potential tensile cracks). If insufficient steel is provided, cracks would develop and
propagate, and could possibly lead to failure.

Reinforcing steel can also supplement concrete in bearing compressive forces, as in columns
provided with longitudinal bars. These bars need to be confined by transverse steel ties [Figure
1-1(d)], in order to maintain their positions and to prevent their lateral buckling. The lateral ties
also serve to confine the concrete, thereby enhancing its compression load-bearing capacity.

The development of reliable design and construction techniques has enabled the construction of a
wide variety of reinforced concrete structures all over the world: building frames (columns and
beams), floor and roof slabs, foundations, bridge decks and piers, retaining walls, grandstands,
water tanks, pipes, bunkers and silos, folded plates and shells, etc.

(a) Plain concrete beam


cracks and fails in
flexural tension under a
small load

(b) Reinforced concrete


beam supports loads
with acceptably low
deformations

CHAPTER 1 - Introduction Page 3


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

(c) Ductile mode of failure


under heavy loads

(d) Reinforced concrete


column

Figure 1-1 - Contribution of steel bars in reinforced concrete

1.3. ADVANTAGES AND DISADVANTAGES OF REINFORCED CONCRETE FOR A


STRUCTURE
The choice of whether a structure should be built of reinforced concrete, steel, masonry, or
timber depends on the availability of materials and on a number of value decisions.

1. Economy.
2. Suitability of material for architectural and structural function.
3. Fire resistance.
4. Rigidity.
5. Low maintenance.
6. Availability of materials.
On the other hand, there are a number of factors that may cause one to select a material other
than reinforced concrete. These include:

1. Low tensile strength.


2. Forms and shoring.

CHAPTER 1 - Introduction Page 4


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

3. Relatively low strength per unit of weight or volume.


4. Time-dependent volume changes.

1.4. THE DESIGN PROCESS

1.4.1. OBJECTIVES OF DESIGN


A structural engineer is a member of a team that works together to design a building, bridge, or
other structure. In the case of a building, an architect generally provides the overall lay-out, and
mechanical, electrical, and structural engineers design individual systems within the building.
The structure should satisfy four major criteria:

1. Appropriateness.
2. Economy.
3. Structural adequacy.
4. Maintainability

1.4.2. THE DESIGN PROCESS


The design process is a sequential and iterative decision-making process. The three major phases
are the following:

1. Definition of the client’s needs and priorities.


2. Development of project concept
3. Design of individual systems.

1.5. DESIGN CODES AND HANDBOOKS

1.5.1. PURPOSE OF CODES


National building codes have been formulated in different countries to lay down guidelines for
the design and construction of structures. The codes have evolved from the collective wisdom of
expert structural engineers, gained over the years. These codes are periodically revised to bring
them in line with current research, and often, current trends.

The codes serve at least four distinct functions:

1. They ensure adequate structural safety, by specifying certain essential minimum


requirements for design.

CHAPTER 1 - Introduction Page 5


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

2. They render the task of the designer relatively simple; often, the results of sophisticated
analyses are made available in the form of a simple formula or chart.
3. The codes ensure a measure of consistency among different designers.
4. They have some legal validity, in that they protect the structural designer from any
liability due to structural failures that are caused by inadequate supervision and/or
faulty material and construction.
The codes are not meant to serve as a substitute for basic understanding and engineering
judgment. The student is, therefore, forewarned that s/he will make a poor designer if s/he
succumbs to the unfortunate (and all-too-common) habit of blindly following the codes. On the
contrary, in order to improve her/his understanding, s/he must learn to question the code
provisions — as, indeed, s/he must, nearly everything in life!

1.5.2. INTRODUCTION TO EUROCODES


The development of the Eurocodes started in 1975; since then they have evolved significantly
and are now claimed to be the most technically advanced structural codes in the world. There are
ten Eurocodes covering all the main structural materials (see Figure 1-2). The structural
Eurocodes were initiated by the European Commission but are now produced by the Comité
Européen de Normalisation (CEN) which is the European standards organization. CEN is
publishing the design standards as full European Standards EN (Euronorms):

EN 1990: Eurocode: Basis of design (EC0)


EN 1991: Eurocode 1 Actions on structures (EC1)
Part 1-1: General actions – Densities, self-weight and imposed loads
Part 1-2: General actions on structures exposed to fire
Part 1-3: General actions – Snow loads
Part 1-4: General actions – Wind loads
Part 1-5: General actions – Thermal actions
Part 1-6: Actions during execution
Part 1-7: Accidental actions from impact and explosions
Part 2: Traffic loads on bridges
Part 3: Actions induced by cranes and machinery

CHAPTER 1 - Introduction Page 6


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Part 4: Actions in silos and tanks


EN 1992: Eurocode 2: Design of concrete structures (EC2)
Part 1-1: General rules and rules for buildings (EC2 Part 1-1)
Part 1-2: General rules - Structural fire design (EC2 Part 1-2)
Part 2: Reinforced and pre-stressed concrete bridges (EC2 Part 2)
Part 3: Liquid retaining and containing structures (EC2 Part 3)
EN 1993: Eurocode 3: Design of steel structures (EC3)
EN 1994: Eurocode 4: Design of composite steel and concrete structures (EC4)
EN 1995: Eurocode 5: Design of timber structures (EC5)
EN 1996: Eurocode 6: Design of masonry structures (EC6)
EN 1997: Eurocode 7: Geotechnical design (EC7)
EN 1998: Eurocode 8: Earthquake resistant design of structures (EC8)
EN 1999: Eurocode 9: Design of aluminum alloy structures (EC9)

All Eurocodes follow a common editorial style. The codes contain ‘Principles’ and ‘Application
rules’. Principles are identified by the letter P following the paragraph number. Principles
are general statements and definitions for which there is no alternative, as well as,
requirements and analytical models for which no alternative is permitted unless specifically
stated.

Application rules are generally recognized rules which comply with the Principles and
satisfy their requirements. Alternative rules may be used provided that compliance with
the Principles can be demonstrated, however the resulting design cannot be claimed to be wholly
in accordance with the Eurocode although it will remain in accordance with Principles.

1. Eurocode: Basis of structural design

In the Eurocode system EN 1990, Eurocode: Basis of Structural Design overarches all the other
Eurocodes (EN 1991 to EN 1999). EN 1990 defines the effects of actions, including geotechnical
and seismic actions, and applies to all structures irrespective of the material of construction. The
material Eurocodes define how the effects of actions are resisted by giving rules for design and
detailing of concrete, steel, composite, timber, masonry and aluminum. (See Figure 1-2).

CHAPTER 1 - Introduction Page 7


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

2. Eurocode 1: Actions on Structures

Eurocode 1 contains in ten parts all the information required by the designer to assess the
individual actions on a structure. It is generally self-explanatory.

Figure 1-2 - The Eurocode Hierarchy

3. Eurocode 2: Design of concrete structures


There are four parts to Eurocode 2;

Eurocode 2, Part 1–1: General rules and rules for buildings, is the principal part which is
referenced by the three other parts.

Eurocode 2, Part 1–2: Structural fire design, gives guidance on design for fire resistance of
concrete structures. Although much of the Eurocode is devoted to fire engineering methods, the
design for fire resistance may still be carried out by referring to tables for minimum cover and
dimensions for various elements.

CHAPTER 1 - Introduction Page 8


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Eurocode 2, Part 2: Bridges, applies the general rules given in Part 1–1 to the design of
concrete bridges. As a consequence both Part 1–1 and Part 2 will be required to carry out a
design of a reinforced concrete bridge.

Eurocode 2, Part 3: Liquid retaining and containment structures, applies the general rules given
in Part 1–1 to the liquid-retaining structures.

1.6. DESIGN PHILOSOPHIES

1.6.1. INTRODUCTION
Over the years, various design philosophies have evolved in different parts of the world, with
regard to reinforced concrete design. A ‘design philosophy’ is built up on a few fundamental
premises (assumptions), and is reflective of a way of thinking.

The earliest codified design philosophy is the working stress method of design (WSM). Close
to a hundred years old, this traditional method of design, based on linear elastic theory, is still
surviving in some countries, although it is now sidelined by the modern limit states design
philosophy.

Historically, the design procedure to follow the WSM was the ultimate load method of design
(ULM), which was developed in the 1950s. Based on the (ultimate) strength of reinforced
concrete at ultimate loads, it evolved and gradually gained acceptance. This method was
introduced as an alternative to WSM in the ACI code in 1956 and the British Code in 1957.

Probabilistic concepts of design developed over the years and received a major impetus from the
mid-1960s onwards. The philosophy was based on the theory that the various uncertainties in
design could be handled more rationally in the mathematical framework of probability theory.
The risk involved in the design was quantified in terms of a probability of failure. Such
probabilistic methods came to be known as reliability-based methods. However, there was
little acceptance for this theory in professional practice, mainly because the theory appeared to
be complicated and intractable (mathematically and numerically).

In order to gain code acceptance, the probabilistic ‘reliability-based’ approach had to be


simplified and reduced to a deterministic format involving multiple (partial) safety factors (rather
than probability of failure). The European Committee for Concrete (CEB) and the International
Federation for Pre-stressing (FIP) were among the earliest to introduce the philosophy of limit

CHAPTER 1 - Introduction Page 9


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

states method (LSM) of design, which is reliability-based in concept. Based on the CEB-FIP
recommendations, LSM was introduced in the British Code CP 110 (1973). In the United States,
LSM was introduced in a slightly different format (strength design and serviceability design) in
the ACI 318−71 (now ACI 318-95).
Thus, the past several decades have witnessed an evolution in design philosophy — from the
traditional ‘working stress method’, through the ‘ultimate load method’, to the modern ‘limit
states method’ of design.

1.6.2. WORKING STRESS METHOD (WSM)


This was the traditional method of design not only for reinforced concrete, but also for structural
steel and timber design. The conceptual basis of WSM is simple. The method basically assumes
that the structural material behaves in a linear elastic manner, and that adequate safety can be
ensured by suitably restricting the stresses in the material induced by the expected ‘working
loads’ (service loads) on the structure. As the specified permissible (‘allowable’) stresses are
kept well below the material strength (i.e., in the initial phase of the stress-strain curve), the
assumption of linear elastic behavior is considered justifiable. The ratio of the strength of the
material to the permissible stress is often referred to as the factor of safety.

The stresses under the applied loads are analyzed by applying the methods of ‘strength of
materials’ such as the simple bending theory. In order to apply such methods to a composite
material like reinforced concrete, strain compatibility (due to bond) is assumed, whereby the
strain in the reinforcing steel is assumed to be equal to that in the adjoining concrete to which it
is bonded. Furthermore, as the stresses in concrete and steel are assumed to be linearly related to
their respective strains, it follows that the stress in steel is linearly related to that in the adjoining
concrete by a constant factor (called the modular ratio), defined as the ratio of the modulus of
elasticity of steel to that of concrete.

However, the main assumption of linear elastic behavior and the tacit assumption that the
stresses under working loads can be kept within the ‘permissible stresses’ are not found to be
realistic. Many factors are responsible for this — such as the long-term effects of creep and
shrinkage, the effects of stress concentrations, and other secondary effects. All such effects
result in significant local increases in and redistribution of the calculated stresses. Moreover,
WSM does not provide a realistic measure of the actual factor of safety underlying a design.

CHAPTER 1 - Introduction Page 10


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

WSM also fails to discriminate between different types of loads that act simultaneously, but have
different degrees of uncertainty. This can, at times, result in very unconservative designs,
particularly when two different loads (say, dead loads and wind loads) have counteracting
effects.

Nevertheless, in defense against these and other shortcomings leveled against WSM, it may be
stated that most structures designed in accordance with WSM have been generally performing
satisfactorily for many years. The design usually results in relatively large sections of structural
members (compared to ULM and LSM), thereby resulting in better serviceability performance
(less deflections, crack-widths, etc.) under the usual working loads. The method is also notable
for its essential simplicity — in concept, as well as application.

1.6.3. ULTIMATE LOAD METHOD (ULM)


With the growing realization of the shortcomings of WSM in reinforced concrete design, and
with increased understanding of the behavior of reinforced concrete at ultimate loads, the
ultimate load method of design (ULM) evolved in the 1950s and became an alternative to WSM.
This method is sometimes also referred to as the load factor method or the ultimate strength
method.

In this method, the stress condition at the state of impending collapse of the structure is analyzed,
and the non-linear stress−strain curves of concrete and steel are made use of. The concept of
‘modular ratio’ and its associated problems are avoided entirely in this method. The safety
measure in the design is introduced by an appropriate choice of the load factor, defined as the
ratio of the ultimate load (design load) to the working load. The ultimate load method makes it
possible for different types of loads to be assigned different load factors under combined loading
conditions, thereby overcoming the related shortcoming of WSM.

This method generally results in more slender sections, and often more economical designs of
beams and columns (compared to WSM), particularly when high strength reinforcing steel and
concrete are used.

However, the satisfactory ‘strength’ performance at ultimate loads does not guarantee
satisfactory ‘serviceability’ performance at the normal service loads. The designs sometimes

CHAPTER 1 - Introduction Page 11


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

result in excessive deflections and crack-widths under service loads, owing to the slender
sections resulting from the use of high strength reinforcing steel and concrete.

1.6.4. LIMIT STATES METHOD (LSM)


The philosophy of the limit states method of design (LSM) represents a definite advancement
over the traditional design philosophies. Unlike WSM, which based calculations on service load
conditions alone, and unlike ULM, which based calculations on ultimate load conditions alone,
LSM aims for a comprehensive and rational solution to the design problem, by considering
safety at ultimate loads and serviceability at working loads.

The LSM philosophy uses a multiple safety factor format which attempts to provide adequate
safety at ultimate loads as well as adequate serviceability at service loads, by considering all
possible ‘limit states’ (defined in the next section). The selection of the various multiple safety
factors is supposed to have a sound probabilistic basis, involving the separate consideration of
different kinds of failure, types of materials and types of loads. In this sense, LSM is more than
a mere extension of WSM and ULM. It represents a new ‘paradigm’ — a modern philosophy.

1. Limit States

When a structure or structural element becomes unfit for its intended use, it is said to have
reached a limit state. The limit states for reinforced concrete structures can be divided into three
basic groups:

I. Ultimate limit states. These involve a structural collapse of part or all of the structure.
Such a limit state should have a very low probability of occurrence, because it may lead
to loss of life and major financial losses. The major ultimate limit states are as follows:
a) Loss of equilibrium of a part or all of the structure as a rigid body. Such a failure would
generally involve tipping or sliding of the entire structure and would occur if the
reactions necessary for equilibrium could not be developed.

b) Rupture of critical parts of the structure, leading to partial or complete collapse. The
majority of this document deals with this limit state. Chapters 3 consider flexural failures;
Chapter 4 shear failures; and so on.

CHAPTER 1 - Introduction Page 12


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

c) Progressive collapse. In some structures, an overload on one member may cause that
member to fail. The load acting on it is transferred to adjacent members which, in turn,
may be overloaded and fail, causing them to shed their load to adjacent members, causing
them to fail one after another, until a major part of the structure has collapsed. This is
called a progressive collapse. Progressive collapse is prevented, or at least is limited, by
one or more of the following:

i. Controlling accidental events by taking measures such as protection against


vehicle collisions or explosions.
ii. Providing local resistance by designing key members to resist accidental events.
iii. Providing minimum horizontal and vertical ties to transfer forces.
iv. Providing alternative lines of support to anchor the tie forces.
v. Limiting the spread of damage by subdividing the building with planes of
weakness sometimes referred to as structural fuses.
A structure is said to have general structural integrity if it is resistant to progressive collapse. For
example, an explosion or a vehicle collision may accidentally remove a column that supports an
interior support of a two-span continuous beam. If properly detailed, the structural system may
change from two spans to one long span. This would entail large deflections and a change in the
load path from beam action to catenary or tension membrane action. Most building Codes
require continuous ties of tensile reinforcement around the perimeter of the building at each floor
to reduce the risk of progressive collapse. The ties provide reactions to anchor the catenary
forces and limit the spread of damage. Because such failures are most apt to occur during
construction, the designer should be aware of the applicable construction loads and procedures.

d) Formation of a plastic mechanism. A mechanism is formed when the reinforcement


yields to form plastic hinges at enough sections to make the structure unstable.
e) Instability due to deformations of the structure. This type of failure involves buckling
f) Fatigue. Fracture of members due to repeated stress cycles of service loads may cause
collapse.
II. Serviceability limit states. These involve disruption of the functional use of the
structure, but not collapse. Because there is less danger of loss of life, a higher
probability of occurrence can generally be tolerated than in the case of an ultimate limit
state. Design for serviceability is discussed in Chapter 4. The major serviceability limit
states include the following:

CHAPTER 1 - Introduction Page 13


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

a) Excessive deflections for normal service. Excessive deflections may cause machinery
to malfunction, may be visually unacceptable, and may lead to damage to nonstructural
elements or to changes in the distribution of forces. In the case of very flexible roofs,
deflections due to the weight of water on the roof may lead to increased depth of water,
increased deflections, and so on, until the strength of the roof is exceeded. This is a
ponding failure and in essence is a collapse brought about by failure to satisfy a
serviceability limit state.
b) Excessive crack widths. Although reinforced concrete must crack before the
reinforcement can function effectively, it is possible to detail the reinforcement to
minimize the crack widths. Excessive crack widths may be unsightly and may allow
leakage through the cracks, corrosion of the reinforcement, and gradual deterioration of
the concrete.
c) Undesirable vibrations. Vertical vibrations of floors or bridges and lateral and
torsional vibrations of tall buildings may disturb the users. Vibration effects have rarely
been a problem in reinforced concrete buildings.
III. Special limit states. This class of limit states involves damage or failure due to abnormal
conditions or abnormal loadings and includes:
a) Damage or collapse in extreme earthquakes,
b) Structural effects of fire, explosions, or vehicular collisions,
c) Structural effects of corrosion or deterioration, and
d) Long-term physical or chemical instability (normally not a problem with
concrete structures).

2. Limit state design process

Limit-states design is a process that involves

1. The identification of all potential modes of failure (i.e., identification of the significant
limit states),
2. The determination of acceptable levels of safety against occurrence of each limit state,
3. Structural design for the significant limit states.
For normal structures, step 2 is carried out by the building-code authorities, who specify the load
combinations and the load factors to be used. For unusual structures, the engineer may need to

CHAPTER 1 - Introduction Page 14


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

check whether the normal levels of safety are adequate. For buildings, a limit-states design starts
by selecting the concrete strength, cement content, cement type, supplementary cementitious
materials, water–cementitious materials ratio, air content, and cover to the reinforcement to
satisfy the durability requirements of Eurocode. Next, the minimum member sizes and minimum
covers are chosen to satisfy the fire-protection requirements of the local building code. Design is
then carried out, starting by proportioning for the ultimate limit states followed by a check of
whether the structure will exceed any of the serviceability limit states. This sequence is followed
because the major function of structural members in buildings is to resist loads without
endangering the occupants. For a water tank, however, the limit state of excessive crack width is
of equal importance to any of the ultimate limit states if the structure is to remain watertight. In
such a structure, the design for the limit state of crack width might be considered before the
ultimate limit states are checked. In the design of support beams for an elevated monorail, the
smoothness of the ride is extremely important, and the limit state of deflection may govern the
design.

1.7. MATERIALS

1.7.1. BEHAVIOR OF CONCRETE UNDER COMPRESSION

1.7.1.1. Compressive strength of concrete


Generally, the term concrete strength is taken to refer to the uniaxial compressive strength as
measured by a compression test of a standard test cylinder, because this test is used to monitor
the concrete strength for quality control or acceptance purposes. For convenience, other strength
parameters, such as tensile or bond strength, are expressed relative to the compressive strength.

1.7.1.2. Statistical Variations in Concrete Strength


Concrete is a mixture of water, cement, aggregate, and air. Variations in the properties or
proportions of these constituents, as well as variations in the transporting, placing, and
compaction of the concrete, lead to variations in the strength of the finished concrete. In addition,
discrepancies in the tests will lead to apparent differences in strength. The shaded area in Figure
1-3 shows the distribution of the strengths in a sample of 176 concrete-strength tests.

CHAPTER 1 - Introduction Page 15


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 1-3 - Distribution of concrete strengths.

The mean or average strength is 3940 psi, but one test has strength as low as 2020 psi and one is
as high as 6090 psi.

If more than about 30 tests are available, the strengths will generally approximate a normal
distribution. The normal distribution curve, shown by the curved line in Figure 1-3, is
symmetrical about the mean value, x of the data. The dispersion of the data can be measured by
the sample standard deviation, S , which is the root-mean-square deviation of the strengths from
their mean value:

( 1-1)

The standard deviation divided by the mean value is called the coefficient of variation, V:

( 1-2)

This makes it possible to express the degree of dispersion on a fractional or percentage basis
rather than an absolute basis. The concrete test data in Figure 1-3 have a standard deviation of
615 psi and a coefficient of variation of or 15.6 percent. 615/3940 = 0.156.

CHAPTER 1 - Introduction Page 16


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

If the data correspond to a normal distribution, their distribution can be predicted from the
properties of such a curve. Thus, 68.3 percent of the data will lie within 1 standard deviation
above or below the mean. Alternatively, 15.6 percent of the data will have values less than
x  s Similarly, for a normal distribution, 10 percent of the data, or 1 test in10, will have

values less than ̅ (1-aV), where a=1.282, Values of a corresponding to other probabilities can
be found in statistics texts.

Figure 1-4 shows the mean concrete strength, fcr , required for various values of the coefficient

of variation if no more than 1 test in 10 is to have strength less than 3000 psi. As shown in this
figure, as the coefficient of variation is reduced, the value of the mean strength, fcr , required to

satisfy this requirement can also be reduced.

Figure 1-4 - Normal frequency curves for coefficients of variation of10, 15, and 20 percent.

NB: Poor control…………….…..V > 14%


Average control…………….V = 10.5%
Excellent control……………V < 7%

1.7.1.3. Factors Affecting Concrete Compressive Strength


 Water/cement ratio.  Moisture conditions during curing.
 Type of cement.  Temperature conditions during curing.

CHAPTER 1 - Introduction Page 17


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

 Supplementary cementitious materials.  Age of concrete


 Aggregate.  Maturity of concrete
 Mixing water.  Rate of loading.

1.7.1.4. Stress-Strain Curves


Typical stress-strain curves of concrete (of various grades), obtained from standard uniaxial
compression tests, are shown in Figure 1-5. The curves are somewhat linear in the very initial
phase of loading; the non-linearity begins to gain significance when the stress level exceeds
about one-third to one-half of the maximum. The maximum stress is reached at a strain
approximately equal to 0.002; beyond this point, an increase in strain is accompanied by a
decrease in stress. For the usual range of concrete strengths, the strain at failure is in the range of
0.003 to 0.005.

The higher the concrete grade, the steeper is the initial portion of the stress-strain curve, the
sharper the peak of the curve, and the less the failure strain. For low-strength concrete, the
curve has a relatively flat top, and a high failure strain.

When the stress level reaches 70–90 percent of the maximum, internal cracks are initiated in the
mortar throughout the concrete mass, roughly parallel to the direction of the applied loading.
The concrete tends to expand laterally, and longitudinal cracks become visible when the lateral
strain (due to the Poisson effect) exceeds the limiting tensile strain of concrete (0.0001—0.0002).
The cracks generally occur at the aggregate-mortar interface. As a result of the associated larger
lateral extensions, the apparent Poisson’s ratio increases sharply.

CHAPTER 1 - Introduction Page 18


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 1-5 - Typical stress-strain curves of concrete in compression

The descending branch of the stress-strain curve can be fully traced only if the strain-controlled
application of the load is properly achieved. For this, the testing machine must be sufficiently
rigid (i.e., it must have a very high value of load per unit deformation); otherwise, the concrete is
likely to fail abruptly (sometimes, explosively) almost immediately after the maximum stress is
reached. The fall in stress with increasing strain is a phenomenon which is not clearly
understood; it is associated with extensive micro-cracking in the mortar, and is sometimes called
softening of concrete.

1.7.1.5. Modulus of Elasticity


The Young’s modulus of elasticity is a constant, defined as the ratio, within the linear elastic
range, of axial stress to axial strain, under uniaxial loading. In the case of concrete under
uniaxial compression, it has some validity in the very initial portion of the stress-strain curve,
which is practically linear [Figure 6]; that is, when the loading is of low intensity, and of very
short duration.

Various descriptions of Ec are possible, such as initial tangent modulus, tangent modulus (at a
specified stress level), secant modulus (at a specified stress level), etc. — as shown in Figure 6.
Among these, the secant modulus at a stress of about one-third the cube strength of concrete is
generally found acceptable in representing an average value of Ec under service load conditions
(static loading).

CHAPTER 1 - Introduction Page 19


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 6 – Various descriptions of modulus of elasticity of concrete


( I T≡ initial tangent, T ≡ tangent, S ≡ secant )

1.7.2. BEHAVIOR OF CONCRETE UNDER TENSION


Concrete is not normally designed to resist direct tension. However, tensile stresses do develop
in concrete members as a result of flexure, shrinkage and temperature changes. Principal
tensile stresses may also result from multi-axial states of stress. Often cracking in concrete is a
result of the tensile strength (or limiting tensile strain) being exceeded. As pure shear causes
tension on diagonal planes, knowledge of the direct tensile strength of concrete is useful for
estimating the shear strength of beams with unreinforced webs, etc. Also, knowledge of the
flexural tensile strength of concrete is necessary for estimation of the ‘moment at first crack’,
required for the computation of deflections and crack widths in flexural members.

As pointed out earlier, concrete is very weak in tension, the direct tensile strength being only
about 7 to 15 percent of the compressive strength. It is difficult to perform a direct tension test
on a concrete specimen, as it requires a purely axial tensile force to be applied, free of any

CHAPTER 1 - Introduction Page 20


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

misalignment and secondary stress in the specimen at the grips of the testing machine. Hence,
indirect tension tests are resorted to, usually the flexure test or the cylinder splitting test.

1.7.2.1. Stress-Strain Curve of Concrete in Tension


Concrete has a low failure strain in uniaxial tension. It is found to be in the range of 0.0001 to
0.0002. The stress-strain curve in tension is generally approximated as a straight line from the
origin to the failure point. The modulus of elasticity in tension is taken to be the same as that in
compression. As the tensile strength of concrete is very low, and often ignored in design, the
tensile stress-strain relation is of little practical value.

1.7.2.2. Splitting Tensile Strength


The cylinder splitting test is the easiest to perform and gives more uniform results compared to
other tension tests. In this test, a ‘standard’ plain concrete cylinder (of the same type as used for
the compression test) is loaded in compression on its side along a diametric plane. Failure
occurs by the splitting of the cylinder along the loaded plane [Figure 7]. In an elastic
homogeneous cylinder, this loading produces a nearly uniform tensile stress across the loaded
plane as shown in Figure 7.

From theory of elasticity concepts, the following formula for the evaluation of the splitting
tensile strength fct is obtained:

2P (1)
fct 
 dL
where P is the maximum applied load, d is the diameter and L the length of the cylinder.

CHAPTER 1 - Introduction Page 21


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 7 – Cylinder splitting test for tensile strength

1.7.3. REINFORCING STEEL


As explained earlier, concrete is reinforced with steel primarily to make up for concrete’s
incapacity for tensile resistance. Steel embedded in concrete, called reinforcing steel, can
effectively take up the tension that is induced due to flexural tension, direct tension, ‘diagonal
tension’ or environmental effects. Reinforcing steel also imparts ductility to a material that is
otherwise brittle. Furthermore, steel is stronger than concrete in compression also; hence,
concrete can be advantageously reinforced with steel for bearing compressive stresses as well, as
is commonly done in columns.

1.7.3.1. Stress-Strain Curves


The stress-strain curve of reinforcing steel is obtained by performing a standard tension test.
Typical stress-strain curves for the three grades of steel are depicted in Figure 1-8.

CHAPTER 1 - Introduction Page 22


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 1-8 - Typical stress-strain curves for reinforcing steels

For all grades, there is an initial linear elastic portion with constant slope, which gives a
modulus of elasticity  Es  that is practically the same for all grades. The Code specifies that the

value of Es to be considered in design is 2 ×105 MPa N/mm2. The stress-strain curve of mild

steel (hot rolled) is characterized by an initial nearly elastic part that is followed by an yield
plateau (where the strain increases at almost constant stress), followed in turn by a strain
hardening range in which the stress once again increases with increasing strain (although at a
decreasing rate) until the peak stress (tensile strength) is reached. Finally, there is a descending
branch wherein the nominal stress (load divided by original area) decreases until fracture occurs.
(The actual stress, in terms of load divided by the current reduced area, will, however, show an
increasing trend).

1.8. EUROCODE’S RECOMMENDATIONS FOR LIMIT STATES DESIGN


The salient features of LSM, as prescribed by the Code, are covered here. Details of the design
procedure for various limit states of collapse and serviceability are covered in subsequent
sections.

CHAPTER 1 - Introduction Page 23


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

1.8.1. ACTIONS
The term action is used in the Eurocodes in order to group together generically all external
influences on a structure’s performance. It encompasses loading by gravity and wind, but
includes also vibration, thermal effects, fire and seismic loading.

Separate combinations of actions are used to check the structure for the design situation being
considered. For each of the particular design situations an appropriate representative value for
each action is used.

1.8.1.1. Representative values of actions


The main actions to be used in load cases used for design are:

 Permanent actions G: e.g. self-weight of structures and fixed equipment;


 Variable actions Q: e.g. imposed loads on building floors and beams; snow loads on
roofs; wind loading on walls and roofs
 Accidental actions A: e.g. fire, explosions and impact.

1.8.1.2. Permanent actions


The characteristic value of a permanent action Gk  may be a single value if variability is known
to be low (e.g. the self-weight of quality-controlled factory-produced members). If the variability
of G cannot be considered as small, and its magnitude may vary from place to place in the
structure, then an upper value Gk ,sup and a lower value Gk ,inf may occasionally be used.

1.8.1.3. Variable actions


Up to four types of representative value may be needed for the variable and accidental actions.
The types most commonly used for variable actions are:

 The characteristic value Qk


and combinations of the characteristic value with other variable actions, multiplied by
different combination factors:
 The combination value  0Qk
 The frequent value  1Qk
 The quasi-permanent value  2Qk
Explanations of the representative values and the design situations in which they arise are given
below. The ‘ x ’ factors generally reduce the value of a variable action present in an accidental
situation compared with the characteristic value.

A. Combination value of  0Qk


The combination value is used for checking:

1. Ultimate limit states;


CHAPTER 1 - Introduction Page 24
AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

2. Irreversible serviceability limit states (e.g. deflections which fracture brittle fittings or
finishes).
It is associated with combinations of actions. The combination factor  0 reduces Qk because of
the low probability of the most unfavourable values of several independent actions occurring
simultaneously.

B. Frequent value  1Qk


The frequent value is used for checking:

1. Ultimate limit states involving accidental actions;


2. Reversible serviceability limit states, primarily associated with frequent combinations.
In both cases the reduction factor  1 multiplies the leading variable action. The frequent value
 1Qk of a variable action Q is determined so that the total proportion of a chosen period of time
during which Q exceeds  1Qk is less than a specified small part of the period.

C. Quasi-permanent value  2Qk


The quasi-permanent value is used for checking:

1. Ultimate limit states involving accidental actions;


2. Reversible serviceability limit states.
Quasi-permanent values are also used for the calculation of long-term effects (e.g. cosmetic
cracking of a slab) and to represent combinations of variable seismic actions. The quasi-
permanent value  2Qk is defined so that the total proportion of a chosen period of time during
which Q exceeds  2Qk is a considerable part (more than half) of the chosen period.

1.8.1.4. Load combinations for design


The values of actions to be used in design are governed by a number of factors. These include:

1. The nature of the load. Whether the action is permanent, variable or accidental, as the
confidence in the description of each will vary.
2. The limit state being considered. Clearly, the value of an action governing design must be
higher for the ultimate limit state than for serviceability for persistent and transient design
situations. Further, under serviceability conditions, loads vary with time, and the design
load to be considered could vary substantially. Realistic serviceability loads should be
modeled appropriate to the aspect of the behavior being checked (e.g. deflection, cracking
or settlement). For example, creep and settlement are functions of permanent loads only.
3. The number of variable loads acting simultaneously. Statistically, it is improbable that all
loads will act at their full characteristic value at the same time. To allow for this, the
characteristic values of actions will need modification.

CHAPTER 1 - Introduction Page 25


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Consider the case of permanent action Gk  and one variable action Qk  only. For the ultimate
limit state the characteristic values should be magnified, and the load may be represented as
 GGk   QQk , where the  factors are the partial safety factors. The values of  G and  Q will be
different, and will be a reflection of the variabilities of the two loads being different. The gamma
factors account for:

1) The possibility of unfavourable deviation of the loads from the characteristic values
2) Inaccuracies in the analyses
3) Unforeseen redistribution of stress
4) Variations in the geometry of the structure and its elements, as this affects the
determination of the action effects.
Now consider the case of a structure subject to variable actions Q1 and Q2 simultaneously. If Q1
and Q2 are independent, i.e. the occurrence and magnitude of Q1 does not depend on the
occurrence and magnitude of Q2 and vice versa, then it would be unrealistic to use
 Q,1Qk ,1   Q,2Qk ,2 as the two loads are unlikely to act at their maximum at the same time. Joint
probabilities will need to be considered to ensure that the probability of occurrence of the two
loads is the same as that of a single load. It will be more reasonable to consider one load at its
maximum in conjunction with a reduced value for the other load. Thus, we have two
possibilities:

 Q,1Qk ,1   0,2  Q,2Qk ,2  (2)

Or

 0,1  Q,1Qk ,1    Q,2Qk ,2 (3)

Multiplication by  0 is said to produce a combination value of the load. It should be noted that
the values of  and  0 vary with each load.

The above discussion illustrates the thinking behind the method of combining loads for an
ultimate limit state check. Similar logic is applied to the estimation of loads for the different
serviceability checks.

I. Ultimate limit state


The following ultimate limit states shall be verified as relevant:

a) EQU: Loss of static equilibrium of the structure or any part of it considered as a rigid
body, where:
 Minor variations in the value or the spatial distribution of actions from a single
source are significant, and
 The strengths of construction materials or ground are generally not governing;

CHAPTER 1 - Introduction Page 26


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

b) STR: Internal failure or excessive deformation of the structure or structural members,


including footings, piles, basement walls, etc., where the strength of construction
materials of the structure governs;
c) GEO: Failure or excessive deformation of the ground where the strengths of soil or rock
are significant in providing resistance;
d) FAT: Fatigue failure of the structure or structural members.
Combinations of actions

1) Persistent and transient situations – fundamental combinations.


In the following paragraphs, various generalized combinations of loads are expressed
symbolically. It should be noted that the ‘+’ symbol in the expressions does not have the normal
mathematical meaning, as the directions of loads could be different. It is best to read it as
meaning ‘combined with’.

EN 1990 gives three separate sets of load combinations, namely EQU (to check against loss of
equilibrium), STR (internal failure of the structure governed by the strength of the construction
materials) and GEO (failure of the ground, where the strength of soil provides the significant
resistance).

Equilibrium: Equilibrium is verified using the load combination Set A in the code, which is as
follows:

 G,JGk , j   Q,1Qk ,1   Q,i 0,i Qk ,i (4)


 G, j ,supGk , j ,sup is used when the permanent loads are unfavourable, and  G, j ,infGk , j ,inf is used when
the permanent actions are favourable. Numerically,  G, j ,sup  1.1 ,  G, j ,inf  0.9 , and  Q  1.5
when unfavourable and 0 when favourable.

The above format applies to the verification of the structure as a rigid body (e.g. overturning of
retaining walls). A separate verification of the limit state of rupture of structural elements should
normally be undertaken using the format given below for strength. In cases where the
verification of equilibrium also involves the resistance of the structural member (e.g.
overhanging cantilevers), the strength verification given below without the above equilibrium
check may be adopted. In such verifications,  G, j ,inf  1.15 should be used.

Strength: when a design does not involve geotechnical actions, the strength of elements should
be verified using load combination Set B. two options are given. Either combination (6.10) from
EN 1990 or the less favourable of equations (6.10a) and (6.10b) may be used:

 G, jGk , j   Q,1Qk ,1   Q,i 0,i Qk ,i (5)

CHAPTER 1 - Introduction Page 27


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

 G, j ,supGk , j ,sup is used when the permanent loads are unfavourable , and  G, j ,infGk , j ,inf is used when
the permanent actions are favourable. Numerically,  G, j ,sup  1.35 ,  G, j ,inf  1.0 , and  Q  1.5
when unfavourable and 0 when favourable (EN1990)

 G, jGk , j   Q,i 0,i Qk ,i (6)


 G, jGk , j   Q,1Qk ,1   Q,i 0,iQk ,i (7)
Numerically,   0.925, G, j ,sup  1.35, G, j ,inf  1.0 and  Q  1.5 when unfavourable and 0 when
favourable (EN 1990)

The above combinations assume that a number of variable actions are present at the same time.
Qk ,1 is the dominant load if it is obvious, otherwise each load is in turn treated as a dominant
load and the other as secondary. The dominant load is then combined with the combination value
of the secondary loads. Both are multiplied by their respective  values.

The magnitude of the load resulting from equations (6.10a) and (6.10b) will always be less than
that from equation (6.10).

Now turning to the factors  G,inf and  G,sup , it will be noted that the numerical values are
different in the verification of equilibrium and that of strength. For instance, in an overhanging
cantilever beam, the multiplier for self-weight in the cantilever section will be 1.1 G,sup  and
that in the anchor span will be 0.9  G,cnf  . The possible explanation for  G,sup being 1.1 and not
1.35 as in the strength check is that

a) The variability in self-weight of the element is unlikely to be large


b) The factor 1.35 has built into it an allowance for structural performance (which is
necessary only for strength checks)
c) The loading in the cantilever will also generally include variable actions, partial safety
factors for which will ensure a reasonable overall safety factor.
When a design involves geotechnical action, a number of approaches are given in EN 1990, and
the choice of the method is a Nationally Determined Parameter.

2) Accidental design situation


The load combination recommended is

Gk , j  Ad   1,iQk ,1   2,iQk ,i (8)


where Ad is the design value of accidental action, Qk ,1 is the main variable action accompanying
the accidental action and Qk ,i are other variable actions.

Accidents are unintended events such as explosions, fire or vehicular impact, which are of short
duration and which have a low probability of occurrence. Also, a degree of damage is generally
CHAPTER 1 - Introduction Page 28
AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

acceptable in the event of an accident. The loading model should attempt to describe the
magnitude of other variable loads which are likely to occur in conjunction with the accidental
load. Accidents generally occur in structures in use. Therefore, the values of variable actions will
be less than those used for the fundamental combination of loads in (1) above. To provide a
realistic variable load combining with the accidental load, the variable actions are multiplied by
different (and generally lower)  factors. Multiplier  1 is applied to the dominant action, and
 2 to the others. Where the dominant action is not obvious, each variable action present is in
turn treated as dominant.  Q for accidental situations is unity.

Multiplication by  1 is said to produce a frequent value of the load, and multiplication by  2 the
quasi-permanent value. Numerical values for  1 and  2 are given in EN 1990.

3) Seismic design situations


Gk , j  AEd   2,iQk ,i (9)
j 1 i 1

II. Serviceability limit state


Combination of actions

1) Characteristic combination.
Gk, j  Qk,1   0,iQk,i (10)
i 1

This represents a combination of service loads, which can be considered rather infrequent. It
might be appropriate for checking sates such as micro cracking or possible local non-catastrophic
failure of reinforcement leading to large cracks in sections.

2) Frequent combination
Gk, j  1,1Qk,1   2,iQk,i i 1 (11)
This represents a combination that is likely to occur relatively frequently in service conditions,
and is used for checking cracking.

3) Quasi-permanent combination
Gk, j   2,iQk,i i  1 (12)
This will provide an estimate of sustained loads on the structure, and will be appropriate for the
verification of creep, settlement, etc.

It should be realized that the above combinations describe the magnitude of loads which are
likely to be present simultaneously. The actual arrangement of loads in position and direction
within the structure to create the most critical effect is a matter of structural analysis (e.g. loading
alternate or adjacent spans in continuous beams).

Values of  factors

CHAPTER 1 - Introduction Page 29


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Table 2 – Recommended values of  factors for buildings

1.8.2. MATERIAL

1.8.2.1. Partial factors for materials


Table 3 – Partial factors for materials for ultimate limit states

The value for partial factors for materials for serviceability limit state verification should be taken as
those given in the particular clauses of this Eurocode. The recommended value is 1.0.

1.8.2.2. Concrete
The compressive strength of concrete is denoted by concrete strength classes which relate to the
characteristic (5%) cylinder strength fck , or the cube strength fck ,cube .

CHAPTER 1 - Introduction Page 30


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

The characteristic strength for fck and the corresponding mechanical characteristics necessary design
are given in the following.

Poisson’s ratio may be taken equal to 0.2 for uncracked concrete and 0 for cracked concrete.

Design compressive strengths

CHAPTER 1 - Introduction Page 31


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

The value of the design compressive strength is defined as

fcd  cc fck  c (13)


Where:

c is the partial safety factor for concrete

 cc is the coefficient taking account of long term effects on the compressive strength and of
unfavourable effects resulting from the way the load is applied.

The value of  cc for use in a Country should lie between 0.8 and 1.0 and may be found in its National
Annex. The recommended value is 1.

Stress-strain relations for the design of cross-sections

For the design of cross-sections, the following stress-strain relationship may be used.

  c  
n (14)
 c  fcd 1   1    for 0   c   c 2
   c 2  
 c  fcd for  c 2   c   cu 2 (15)
Where:

n is the exponent according to Table 3.1 of Eurocode


c2 is the strain at reaching the maximum strength according to Table 3.1 of Eurocode
 cu is the ultimate strain according to Table 3.1 of Eurocode

Figure 9 – Parabola-rectangle diagram for concrete under compression

CHAPTER 1 - Introduction Page 32


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Other simplified stress-strain relationships may be used if equivalent to or more conservative than the
one defined above, for instance bi-linear according to the following figure (compressive stress and
shortening strain shown as absolute values) with values of  c 3 and  cu 3 .

Figure 10 – Bi-Linear stress-strain relation

A rectangular stress distribution as given in the figure below may be assumed. The factor  , defining
the effective height of the compression zone and the factor  , defining the effective strength, follow
from:

  0.8 for fck  50 MPa (16)


  0.8   fck  50 400 for 50  fck  90 MPa (17)
and

  1.0 for fck  50 MPa (18)


  1.0   fck  50 200 for 50  fck  90 MPa (19)

CHAPTER 1 - Introduction Page 33


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 11 – Rectangular stress distribution

1.8.2.3. Reinforcing steel


The application rules for design and detailing in this Eurocode are valid up for specified yield strength,
fyk  400  600 MPa.

For normal design, either of the following assumptions may be made

a. an inclined top branch with a strain limit of  ud and a maximum stress of kfyk  s at  uk ,


where k  ft fy  k

b. A horizontal top branch without the need to check the strain limit.

The recommended value of  ud is 0.9 uk and the value of ft fy  k
is given in Annex C of Eurocode 2.

Figure 12 – Idealized and design stress-strain diagrams for reinforcing steel (for tension and
compression)

CHAPTER 1 - Introduction Page 34


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

The mean value of density may be assumed to be 7850 kg/m3.The design value of the modulus of
elasticity Es may be assumed to be 200 GPa.

CHAPTER 1 - Introduction Page 35


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

CHAPTER 2. LIMIT STATE DESIGN FOR FLEXURE

2.1. INTRODUCTION

2.1.1. ANALYSIS VS. DESIGN


Two different types of problems arise in the study of reinforced concrete:

1 Analysis – Given a cross section, concrete strength, reinforcement size and location, and
yield strength, compute the resistance or strength. In analysis there should be one unique
answer.
2 Design – Given a factored design moment, select a suitable cross section, including
dimensions, concrete strength, reinforcement, and so on. In design there are many possible
solutions.

2.1.2. STATICS OF BEAM ACTION


A beam is a structural member that supports applied loads and its own weight primarily by
internal moments and shears. Figure 2-1a shows a simple beam that supports its own dead
weight,  per unit length, plus a concentrated load, P. if the axial applied load, N, is equal to
zero, as shown, the member is referred to as a beam. If N is a compressive force, the member is
called a beam-column. This chapter will be restricted to the very common case where N  0 .

Chapter 2 – Limit State Design for Flexure Page 36


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 2-1 – Internal forces in a beam

The loads  and P cause bending moments, distributed as shown in Figure 2-1b. The bending
moments is a load effect calculated from the loads by using the laws of statics. For a simply
supported beam of a given span and for a given set of loads  and P, the moments are
independent of the composition and size of the beam.

At any section within the beam, the internal resisting moment, M, shown in Figure 2-1c is
necessary to equilibrate the bending moment. An internal resisting shear, V, also is required, as
shown.

The internal resisting moment, M, results from an internal compressive force, C, and an internal
tensile force, T, separated by a lever arm, jd, as shown in Figure 2-1d. Because there are no
external axial loads, summation of the horizontal forces gives

C  T  0 or C  T (2-1)
If moments are summed about an axis through the point of application of the compressive force,
C, the moment equilibrium of the free body gives

M  T  jd (2-2)
Similarly, if moments are summed about the point of application of the tensile force, T,

M  C  jd (2-3)
Because C = T, these two equations are identical. Equations (2-1), (2-2) and (2-3) come directly
from statics and are equally applicable to beams made of steel, wood, or reinforced concrete.

The conventional elastic beam theory results in the equation   My I , which, for an
uncracked, homogeneous rectangular beam without reinforcement, gives the distribution of
stresses shown in Figure 2-2.

Chapter 2 – Limit State Design for Flexure Page 37


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 2-2 – Elastic beam stresses and stress blocks

The stress diagram shown in Figure 2-2c and Figure 2-2d may be visualized as having a
“volume”; hence, one frequently refers to the compressive stress block. The resultant
compressive force C, which is equal to the volume of the compressive stress block in Figure
2-2d, is given by

 c max   h  (2-4)
C b 2 
2  

In a similar manner, one could compute the force T from the tensile stress block. The forces C
and T act through the centroids of the volumes of the respective stress blocks. In the elastic case,
these forces act at h 3 above or below the neutral axis, so that jd  2h 3 .

From Equations (2-3)and (2-4) and Figure 2-2, we can write

bh  2h  (2-5)
M   c (max)
4  3 
bh3 12 (2-6)
M   c (max)
h2
or, because

bh3 (2-7)
I
12
And

Chapter 2 – Limit State Design for Flexure Page 38


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

ymax  h 2 (2-8)
It follows that

 c (max)I (2-9)
M
y max
Thus, for the elastic case, identical answers are obtained from the traditional beam stress
equation (2-9), and when the stress block concept is used in equation (2-5)

The elastic beam theory in equation (2-9) is not used in the design of reinforced concrete beams,
because the compressive stress-strain relationship for concrete becomes nonlinear at higher strain
values. What is even more important is that concrete cracks at low tensile stresses, making it
necessary to provide steel reinforcement to carry the tensile force, T.

2.2. DISTRIBUTION OF STRAINS AND STRESSES ACROSS A SECTION IN


BENDING
The theory of bending for reinforced concrete assumes that the concrete will crack in the regions
of tensile strains and that, after cracking, all the tension is carried by the reinforcement. It is also
assumed that plane sections of a structural member remain plane after straining, so that across
the section there must be a linear distribution of strains.

Figure 2-3 – Singly reinforced rectangular beam

Figure 2-3 shows the cross-section of a member subjected to bending, and the resultant strain
diagram, together with three different types of stress distribution in the concrete:

1. The triangular stress distribution applies when the stresses are very nearly proportional to
the strains, which generally occurs at the loading levels encountered under working
conditions and is, therefore, used at the serviceability limit state.
2. The rectangular-parabolic stress block represents the distribution at failure when the
compressive strains are within the plastic range, and it is associated with the design for
the ultimate limit state.

Chapter 2 – Limit State Design for Flexure Page 39


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

3. The equivalent rectangular stress block is a simplified alternative to the rectangular


parabolic distribution.

2.3. ULTIMATE LIMIT STATE FOR FLEXURE

2.3.1. BASIC ASSUMPTIONS FOR FLEXURE AT THE ULS


The theory of flexure for reinforced concrete is based on three basic assumptions, which are
sufficient to allow one to calculate the moment resistance of a beam.

1. Sections perpendicular to the axis of bending that are plane before bending remain plane
after bending.
2. The strain in the reinforcement is equal to the strain in the concrete at the same level.
3. The stresses in the concrete and reinforcement can be computed from the strains by using
stress-strain curves for concrete and steel.
4. The tensile strength of the concrete is ignored.

The first of these is the traditional “plane sections remain plane” assumption made in the
development of flexural theory for beams constructed with any material. The second assumption
is necessary, because the concrete and the reinforcement must act together to carry load. This
assumption implies a prefect bond between the concrete and the steel.

However, the assumptions are not strictly true. The deformations within a section are very
complex, and, locally, plane sections do not remain plane. Nor, due to local bond slip, are the
strains in the concrete exactly the same as those in the steel. Nevertheless, on average, the
assumptions are correct, and are certainly sufficiently true for practical purposes for design of
normal members.

2.3.2. POSSIBLE RANGE OF STRAIN DISTRIBUTIONS AT ULS


The possible range of strain distributions given in EN 1992-1-1-2004 is shown in Figure 2-4.

Chapter 2 – Limit State Design for Flexure Page 40


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 2-4 – Possible strain distributions in the ultimate limit state

A more elaborative diagram for the possible strain distributions is shown in the figure below

Figure 2-5 – Ultimate Strain distributions

2.3.3. LIMITING COMPRESSIVE STRAINS AT ULS


It is universal to define failure of concrete in compression by means of a limiting compressive
strain. The formulation of the limit varies from code to code, for example the American Concrete
Institute code, ACI 318, uses a limit of 0.003, while the UK code BS 8110 uses 0.0035. For
concrete strengths not exceeding 50 N/mm2, the Eurocode adopts values of 0.0035 for flexure
Chapter 2 – Limit State Design for Flexure Page 41
AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

and for combined bending and axial load where the neutral axis remains within the section, and a
limit of between 0.0035 and 0.002 for sections loaded so that the whole section is in
compression.

The logic behind the reduction in the strain limit for axial compression is that, in axial
compression, failure will occur at the strain corresponding to the attainment of the maximum
compressive stress. This is 0.002 for concrete strengths not exceeding 50 N/mm2. In flexure,
considerably higher strains can be reached before the maximum capacity of the section is
reached, and the value of 0.0035 has been obtained empirically.

2.4. TYPES OF FLEXURAL FAILURES


There are three types of flexural failures of reinforced concrete sections: tension, compression
and balanced failures. These three types of failures may be discussed to choose the desirable type
of failure from the three, in case failure is imminent.

A. Tension Failure

If the steel content As of the section is small, the steel will reach fyd before the concrete reaches
its maximum strain of εcu . With further increase in loading, the steel force remains constant at fyd
As, but results a large plastic deformation in the steel, wide cracking in the concrete and large
increase in compressive strain in the extreme fiber of concrete. With this increase in strain the
stress distribution in the concrete becomes distinctly non-linear resulting in increase of the mean
stress. Because equilibrium of internal forces should be maintained, the depth of the N.A
decreases, which results in the increment of the lever arm z. The flexural strength is reached
when concrete strain reaches εcu . This phenomenon is shown in Figure 2-6. This type of failure
is preferable and is used for design.

Figure 2-6 – Tension Failure

B. Compression Failure

Chapter 2 – Limit State Design for Flexure Page 42


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

If the steel content As is large, the concrete may reach its capacity before steel yields. In such a
case the N.A depth increases considerably causing an increase in compressive force. Again the
flexural strength of the section is reached εcu. The section fails suddenly in a brittle fashion. This
phenomenon is shown in Figure 2-7.

Figure 2-7 – Compression Failure

C. Balanced Failure

At balanced failure the steel reaches fyd and the concrete reaches a strain of εcu simultaneously.
This phenomenon is shown in Figure 2-8.

Figure 2-8 – Balanced Failure

2.5. ANALYSIS OF BEAMS FOR FLEXURE AT THE ULS


Two requirements are satisfied throughout the flexural analysis and design of reinforced concrete
beams and columns:

Chapter 2 – Limit State Design for Flexure Page 43


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

1. Stress and strain compatibility: The stress at any point in a member must correspond to
the strain at that point
2. Equilibrium: Internal forces must balance the external load effects

2.5.1. ANALYSIS OF SINGLY REINFORCED BEAM SECTIONS


The general procedure of analysis of singly reinforced concrete beams for its flexural
resistance according to EN 1992-1-1-2004 is as follows.

 Step 1: Assume the type of failure

From section 2.4, there are three possible types of failure for reinforced
concrete beams under flexure. These are compression failure, tension failure
and balanced failure.

 Step 2: Draw the strain profile corresponding to the type of failure

a. Tension failure b. Compression failure c. Balanced failure

 Step 3: Take any of the three possible stress strain relationships for concrete described
in chapter 1 to define the stress block

a. Using parabola-rectangle diagram

Chapter 2 – Limit State Design for Flexure Page 44


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

b. Using Bi-linear diagram

c. Using rectangular diagram

 Step 4: Take the stress strain relationship for the reinforcement bar

 Step 5: Apply condition of equilibrium to the given stress block and conditions of
compatibility to the strain profile to estimate the neutral axis depth

 Step 6: Calculate the strain in the reinforcement bar and check if the assumed type of
failure is correct

 Step 7: If the assumption is correct, apply the moment equilibrium to the stress block
and estimate the moment capacity

 Step 8: If it is not correct, assume another type of failure and repeat steps 2 to step 6
until the assumption is proven to be true

2.5.2. ANALYSIS OF DOUBLY REINFORCED BEAM SECTIONS


Occasionally, beam sections are designed to have both tension reinforcement and
compression reinforcement. These are referred to as doubly reinforced sections.

Chapter 2 – Limit State Design for Flexure Page 45


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

The general procedure of analysis of doubly reinforced concrete beams for its flexural
resistance according to EN 1992-1-1-2004 is as follows.

 Step 1: Assume the type of failure

 Step 2: Draw the strain profile corresponding to the type of failure

 Step 3: Assume the strain in the negative reinforcement either to be greater than the
yield strain or to be less than the yield strain

 Step 4: Take any of the three possible stress strain relationships for concrete
described in chapter 1 to define the stress block

 Step 5: Take the stress strain relationship for the reinforcement bar

 Step 6: Apply condition of equilibrium to the given stress block and conditions of
compatibility to the strain profile to estimate the neutral axis depth

 Step 7: Calculate the strain in the negative reinforcement bars and check if the
assumption is step 3 is correct.

 Step 8: If the assumption is true, proceed to step 8, otherwise revise the assumption in
step 3 and repeat steps 4 to 7.

 Step 9: Calculate the strain in the positive reinforcement bar and check if the assumed
type of failure is correct

 Step 10: If the assumption is correct, apply the moment equilibrium to the stress block
and estimate the moment capacity

 Step 11: If it is not correct, assume another type of failure and repeat steps 2 to step 6
until the assumption is proven to be true

2.5.3. ANALYSIS OF FLANGED SECTIONS

2.5.3.1. Introduction
In the floor system shown in Figure 2-9, the slab is assumed to carry the loads in one direction to
beams that carry them in the perpendicular direction. During construction, the concrete in the
columns is placed and allowed to harden before the concrete in the floor is placed. In the next
construction operation, concrete is placed in the beams and slab in a monolithic pour. As a result,
the slab serves as the top flange of the beams, as indicated by the shading in Figure 2-9. Such a
beam is referred to as a T-beam. The interior beam, AB, has flange on both sides. The spandrel
beam, CD, with a flange on one side only, is often referred to as an inverted L-beam.

Chapter 2 – Limit State Design for Flexure Page 46


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 2-9 – T-beams in a one-way beam and slab floor

Figure 2-10 – Positive and negative moment regions in a T-beam

An exaggerated deflected view of the interior beam is shown in Figure 2-10. This beam develops
positive moments at midspan (section A-A) and negative moments over the supports (section B-
B). At midspan, the compression zone is in the flange, as shown in Figure 2-10b and Figure
2-10d. Generally, it is rectangular, as shown Figure 2-10b, although, in very rare cases for typical
Chapter 2 – Limit State Design for Flexure Page 47
AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

reinforced concrete construction, the neutral axis may shift down into the web, giving a T-shaped
compression zone, as shown in Figure 2-10d. At the support, the compression zone is at the
bottom of the beam and is rectangular, as shown in Figure 2-10c.

2.5.3.2. Effective flange width


The forces acting on the flange of a simply supported T-beam are illustrated in Figure 2-11. At
the support, there are no longitudinal compressive stresses in the flange, but at midspan, the full
width is stressed in compression. The transition requires horizontal shear stresses on the web-
flange interface as shown in Figure 2-11. As a result there is a “shear-lag” effect, and the
portions of the flange closest to the web are more highly stressed than those portions of the
flange closest to the web are more highly stressed than those portions farther away, as shown in
Figure 2-11 and Figure 2-12.

Figure 2-11 – Actual flow of forces on a T-beam flange

Figure 2-12a shows the distribution of the flexural compressive stresses in a slab that forms the
flanges of a series of parallel beam at a section of maximum positive moment. The compressive
stress is a maximum over each web, dropping between the webs. When analyzing and designing
the section for positive moments, an effective compression flange width is used (Figure 2-12b).
When this width, be , is stressed uniformly, it will give approximately the same compression
force that actually is developed in the full width of the compression zone.

Chapter 2 – Limit State Design for Flexure Page 48


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 2-12 – Effective width of T-beams

According to EN 1992-1-1-2004, In T beams the effective flange width, over which uniform
conditions of stress can be assumed, depends on the web and flange dimensions, the type of
loading, the span, the support conditions and the transverse reinforcement.

The effective width of flange should be based on the distance l 0 between points of zero moment,
which may be obtained from Figure 2-13.

Figure 2-13 – Definition of I0 , for calculation of flange width

Note: The length of the cantilever, l 3 , should be less than half the adjacent span and the ration of
adjacent spans should lie between 2/3 and 1.5.

The effective flange width parameters are shown in Figure 2-12 below.

Chapter 2 – Limit State Design for Flexure Page 49


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 2-14 – Effective flange width parameters

The effective flange width beff for a T beam or L beam may be derived as:

beff   beff ,i  bw  b (2-10)


where

beff ,i  0.2bi  0.1l0  0.2l0 (2-11)


and

beff ,i  bi (2-12)

2.5.3.3. Procedure of analysis of flanged beam for flexure


a) Flanged beam subjected to negative moment

For a flanged beam with a negative moment, the compression zone will be the bottom
rectangular part of the web, thus following the procedures for analysis of rectangular sections
will be appropriate.

b) Flanged beam subjected to positive moment

If a flanged beam is subjected to positive moment, the neutral axis might remain within the
flange of the beam or it might be in the web of the beam.

For the case where the neutral axis remains in the flange, the section may be treated as a
rectangular section, and the procedures of analysis of rectangular sections can be adopted.
However, if the neutral axis is in the web of the beam, a different approach for analysis is
necessary and in doing so, adopting the rectangular stress relationship for the concrete in
compression will simplify the analysis.

The general procedure for the analysis of flanged beam subjected to positive moment according
to EN 1992-1-1-2004 is as follows.

Chapter 2 – Limit State Design for Flexure Page 50


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

 Step 1: Assume the neutral axis to be in the flange

 Step 2: Use the procedure of analysis of singly reinforced concrete sections to


estimate neutral axis depth

 Step 3: Check if the assumption is step 1 is correct

 Step 4: If the assumption is correct, estimate the moment resistance of the section
using the procedures of singly reinforced concrete sections. If not correct,
proceed to step 5.

 Step 5: Assume the strain in the tension reinforcement to be greater than the yield
strain.

 Step 6: Take the rectangular stress strain relationship for the concrete under
compression

 Step 7: Take the stress strain relationship for the reinforcement bar

 Step 8: Apply condition of equilibrium to the given stress block and conditions of
compatibility to the strain profile to estimate the neutral axis depth

 Step 9: Calculate the strain in the reinforcement bar and check if the assumed type of
failure is correct

 Step 10: If the assumption is correct, apply the moment equilibrium to the stress block
and estimate the moment capacity

 Step 11: If it is not correct, assume another type of failure and repeat steps 6 to step 10
until the assumption is proven to be true

2.6. DESIGN OF BEAMS FOR FLEXURE AT ULS

2.6.1. CONCRETE COVER


It is necessary to have cover (concrete between the surface of the slab or beam and the
reinforcing bars) for four primary reasons:

1. To bond the reinforcement to the concrete so that the two elements act together. The
efficiency of the bond increases as the cover increases.
2. To protect the reinforcement against corrosion.
3. To protect the reinforcement from strength loss die to overheating in the case of fire.
4. Additional cover sometimes is provided on the top of slabs, particularly in garages and
factories, so that abrasion and wear due to traffic will not reduce the cover below that
required for structural and other purposes.

Chapter 2 – Limit State Design for Flexure Page 51


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

The concrete cover is the distance between the surface of the reinforcement closest to the nearest
concrete surface (including links and stirrups and surface reinforcement where relevant) and the
nearest concrete surface.

Concrete cover according to EN 1992-1-1 and EN 1992-1-2

The nominal cover is defined as a minimum cover, cmin , plus an allowance in design for
deviation, cdev :

cnom  cmin  cdev ( 2-13 )


where cmin should be set to satisfy the requirements below:

 Safe transmission of bond forces


 Durability
 Fire resistance

and cdev is an allowance which should be made in the design for deviations from the minimum
cover. It should be taken as 10 mm, unless fabrication (i.e. construction) is subjected to a quality
assurance system, in which case it is permitted to reduce cdev to 5 mm.

cmin  max cmin,b ; cmin,dur ;10 mm

1. Minimum cover for bond, cmin,b

The minimum cover to ensure adequate bond should not be less than the bar diameter, unless the
aggregate size is over 32 mm. if the aggregate size is over 32 mm, cmin,b should be increased by 5
mm.

2. Minimum cover for durability

EC-2 leaves the choice of Cmin,dur to countries, but gives the following recommendation:

The value of Cmin,dur depends on the “structural class”, which has to be determined first. If the
specified service life is 50 years, the structural class is defined as 4. The “structural class” can be
modified in case of the following conditions:

 The service life is 100 years instead of 50 years


 The concrete strength is higher than necessary
 Slabs (position of reinforcement not affected by construction process)
 Special quality control measures apply

Chapter 2 – Limit State Design for Flexure Page 52


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

The finally applying service class can be calculated with Table 4.3 N but the recommended
minimum structural class is 1.

Chapter 2 – Limit State Design for Flexure Page 53


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

3. Minimum cover for fire resistance

Rather than giving the minimum cover, the tabular method is based on nominal axis distance, see
fig. this is the distance from the center of the main reinforcing bar to the surface of the member.
The designer should ensure that

a  cnom  link  bar 2 ( 2-14 )

Chapter 2 – Limit State Design for Flexure Page 54


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Table 2-1 – Minimum dimensions and axis distances for simply supported beams made with
reinforced and prestressed concrete

Chapter 2 – Limit State Design for Flexure Page 55


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Table 2-2 – Minimum dimensions and axis distances for continuous beams made with reinforced
and prestressed concrete

2.6.2. MINIMUM AND MAXIMUM AREA OF REINFORCEMENT


Minimum and Maximum area of reinforcement according to EN 1992-1-1-2004

The area of longitudinal tension reinforcement should not be taken as less than As,min

fctm ( 2-15 )
As,min  0.26 bt d but not less than 0.0013bt d
fyk
Where:

Chapter 2 – Limit State Design for Flexure Page 56


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

bt denotes the mean width of the tension zone; for a T-beam with the flange in
compression, only the width of the web is taken into account in calculating the
value of bt .
fctm should be determined with respect to the relevant strength class according to Table
3.1. of Eurocode
The cross-sectional area of tension or compression reinforcement should not exceed As,max
outside lap locations.

The value of As,max for beams for use in a country may be found in its National Annex. The
recommended value is 0.04 Ac .

2.6.3. DESIGN OF SINGLY REINFORCED BEAM SECTIONS

2.6.3.1. Limit to the use of singly reinforced sections


At the ultimate limit state it is important that member sections in flexure should be ductile and
that failure should occur with the gradual yielding of the tension steel and not by a sudden
catastrophic compression failure of the concrete. Also, yielding of the reinforcement enables the
formation of plastic hinges so that redistribution of maximum moments can occur, resulting in a
safer and more economical structure.

To ensure rotation of the plastic hinges with sufficient yielding of the tension steel and also to
allow for other factors such as the strain hardening of the steel, Clause 5.5 in EN 1992-1-1 give
limits to the neutral axis depth at the ultimate limit state as a function of the amount of
redistribution carried out in the analysis.

x d    0.4   0.6  0.0014  cu  ( 2-16)


In applying this formula, some degree of interpretation of the code is necessary since the reader
is referred to Table 3.1 of EN 1992-1-1 for values of the ultimate strain. Table 3.1 gives several
values, and it has been assumed that the ultimate strain is  cu1 in this table. x/d should not be
greater than 0.45 for concrete grades C50 or below, and not greater than 0.35 for grades C55 or
higher.

 in equation ( 2-16) is the ratio of the redistributed moment to the moment before
redistribution.  is limited as a function of the type of reinforcement used as follows: for Class
B and Class C steel,   0.70 and for Class A steel,   0.8 .

2.6.3.2. Design procedure


The general procedure for the design of singly reinforced beams according to EN 1992-1-1 is as
follows.

 Step 1: Draw a the strain profile that results a ductile failure by setting the ultimate
strain in the concrete
Chapter 2 – Limit State Design for Flexure Page 57
AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

 Step 2: Take any of the three possible stress strain relationships for concrete described
in chapter 1 to define the stress block

a. Using parabola-rectangle diagram

b. Using Bi-linear diagram

c. Using rectangular diagram

 Step 3: Apply force equilibrium and moment equilibrium to get the neutral axis depth

 Step 4: If the assumption is correct, estimate the moment resistance of the section
using the procedures of singly reinforced concrete sections. If not correct,

Chapter 2 – Limit State Design for Flexure Page 58


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

proceed to step 5.

 Step 5: Check if the ratio of neutral axis depth to the effective depth is below the limit
to the use of singly reinforced sections

 Step 6: If it is below the limit, calculate the area of reinforcement

 Step 7: If x/d exceeds the limit go to section and design the beam as double reinforced

 Step 8: Check the minimum and maximum area of reinforcement

2.6.4. DESIGN OF DOUBLY REINFORCED SECTIONS


The general procedure for the design of doubly reinforced beams according to EN 1992-1-1 is as
follows.

 Step 1: Draw the strain profile that results a ductile failure by setting the ultimate strain
in the concrete

 Step 2: Take any of the three possible stress strain relationships for concrete described
in chapter 1 to define the stress block

 Step 3: Apply force and moment equilibrium to get the neutral axis depth and the area
of reinforcement

 Step 4: Check the minimum and maximum area of reinforcement

2.6.5. DESIGN OF FLANGED BEAMS


a) Flanged beam subjected to negative moment

For a flanged beam with a negative moment, the compression zone will be the bottom
rectangular part of the web, thus following the procedures for design of rectangular sections will
be appropriate.

b) Flanged beam subjected to positive moment

If a flanged beam is subjected to positive moment, the neutral axis might remain within the
flange of the beam or it might be in the web of the beam.

For the case where the neutral axis remains in the flange, the section may be treated as a
rectangular section with width b as the width of the flange, and the procedures of analysis of
rectangular sections can be adopted. However, if the neutral axis is in the web of the beam, a
different approach for design is necessary and in doing so, adopting the rectangular stress
relationship for the concrete in compression will simplify the procedure.

Chapter 2 – Limit State Design for Flexure Page 59


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

The general procedure for the design of flanged beam subjected to positive moment according to
EN 1992-1-1 is as follows.

 Step 1: Assume the neutral axis to be in the flange of the section

 Step 2: Draw the strain profile that results a ductile failure by setting the ultimate
strain in the concrete

 Step 3: Take the simplified rectangular stress block for concrete

 Step 4: Apply force equilibrium and moment equilibrium to get the neutral axis depth

 Step 5: Check if the assumption in step 1 is correct

 Step 6: If the assumption is correct, calculate the necessary area of reinforcement, if


not, proceed to step 7

 Step 7: Take the neutral axis to be in the web

 Step 8: Use force and moment equilibrium to get the necessary area of reinforcement

 Step 9: Check the minimum and maximum area of reinforcement

Chapter 2 – Limit State Design for Flexure Page 60


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

CHAPTER 3. LIMIT STATE DESIGN FOR SHEAR

3.1. THEORETICAL BACKGROUND

3.1.1. DIAGONAL TENSION IN HOMOGENEOUS ELASTIC BEAMS


The stresses acting in homogenous beams can be derived from mechanics of elastic materials.
Shear stresses

VQ (20)

Ib

act at any section in addition to the bending stresses

My (21)

I

except for those locations at which the shear force V happens to be zero.

The role of shear stresses is easily visualized by the performance under load of the laminated
beam of Figure 3-1; it consists of two rectangular pieces bonded together along the contact
surface. If the adhesive is strong enough, the member will deform as one single beam, as shown
in Figure 3-1a . On the other hand, if the adhesive is weak, the two pieces will separate and slide
relative to each other, as shown in Figure 3-1b .

Figure 3-1 – Shear in homogeneous rectangular beams

CHAPTER 3 – Limit State Design for Shear Page 61


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Evidently, then, when the adhesive is effective, there are forces or stresses acting in it that
prevent this sliding or shearing. These horizontal shear stresses are shown in Figure 3-1c as they
act, separately, on the top and bottom pieces. The same stresses occur in horizontal planes in
single-piece beams; they are different in intensity at different distances from the neutral axis.

Figure 3-1d shows a differential length of a single-piece rectangular beam acted upon by a shear
force of magnitude V. Upward translation is prevented; i.e., vertical equilibrium is provided by
the vertical shear stresses  . Their average value is equal to the shear force divided by the cross-
sectional area  av  V ab , but their intensity varies over the depth of the section. The shear
stress is zero at the outer fibers and has a maximum of 1.5 av at the neutral axis, the variation
being parabolic. If a small square element located at the neutral axis of such a beam is isolated as
shown in Figure 3-2b, the vertical shear stresses on it, equal and opposite on the two faces for
reasons of equilibrium, act as shown. However, if these were the only stresses present, the
element would not be in equilibrium; it would spin. Therefore, on the two horizontal faces there
exist equilibrating horizontal shear stresses of the same magnitude. That is, at any point within
the beam, the horizontal shear stresses of Figure 3-2b are equal in magnitude to the vertical shear
stresses of Figure 3-1d.

Figure 3-2 – Stress trajectories in homogeneous rectangular beam

CHAPTER 3 – Limit State Design for Shear Page 62


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

It is proved in any strength-of-materials text that on an element cut at 450 these shear stresses
combine in such a manner that their effect is as shown in Figure 3-2c. That is, the action of the
two pairs of shear stresses on the vertical and horizontal faces is the same as that of two pairs of
normal stresses, one tensile and one compressive, acting on the 450 faces and of numerical value
equal to that of the shear stresses. If an element of the beam is considered that is located neither
at the neutral axis nor at the outer edges, its vertical faces are subject not only to the shear
stresses but also to the familiar bending stresses. The six stresses that now act on the element can
again be combined into a pair of inclined compressive stresses and a pair of inclined tensile
stresses that act at right angles to each other. They are known as principal stresses (Figure 3-2e).

Since the magnitudes of the shear stresses  and the bending stresses f change both along the
beam and vertically with distance from the neutral axis, the inclinations as well as the
magnitudes of the resulting principal stresses also vary from one place to another. Figure 3-2f
shows the inclinations of these principal stresses for a uniformly loaded rectangular beam. That
is, these stress trajectories are lines which, at any point, are drawn in that direction in which the
particular principal stress, tension or compression, acts at that point. It is seen that at the neutral
axis the principal stresses in a beam are always inclined at 450 to the axis. In the vicinity of the
outer fibers they are horizontal near midspan.

An important point follows from this discussion. Tensile stresses, which are of particular concern
in view of the low tensile strength of the concrete, are not confined to the horizontal bending
stresses f that are caused by bending alone. Tensile stresses of various inclinations and
magnitudes, resulting from shear alone (at the neutral axis) or from the combined action of shear
and bending, exist in all parts of a beam and can impair its integrity if not adequately provided
for. It is for this reason that the inclined tensile stresses, known as diagonal tension, must be
carefully considered in reinforced concrete design.

3.1.2. BEHAVIOR OF BEAMS FAILING IN SHEAR

3.1.2.1. Behavior of beams without web reinforcement


The forces transferring shear across an inclined crack in a beam without web reinforcements are
illustrated in Figure 3-3.

CHAPTER 3 – Limit State Design for Shear Page 63


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 3-3 – Internal forces in a cracked beam without web reinforcements

Shear is transferred across line A-B-C by v cy , the shear in the compression zone, by v ay , the
vertical component of the shear transferred across the crack by interlock of the aggregate
particles on the two faces of the crack, and by v d , the dowel action of the longitudinal
reinforcement. Immediately after inclined cracking, as much as 40 to 60 percent of the total shear
is carried by v d and v ay together.

Considering D-E-F portion of the beam below the crack and summing moments about the
reinforcement at point E shows that v d and v a cause a moment about E that must be
equilibrated by a compression force C1 . Horizontal force equilibrium on section A-B-D-E
shows that T1  C1  C1 , and finally, T1 and C1  C1 must equilibrate the external moment at
this section.

As the crack widens, v a decreases, increasing the fraction of the shear resisted by v cy and v d .
The dowel shear, v d , leads to a splitting crack in the concrete along the reinforcement. When
 and C1 ,
this crack occurs, v d drops, approaching zero. When v a and v d disappear, so do v cy
with the result that all the shear and compression are transmitted in the depth AB above the
crack. At this point in the life of the beam, the section A-B is too shallow to resist the
compression forces needed for equilibrium. As a result, this region crushes or buckles upward.

Note also that if C1  0 , then T2  T1 , and as a result, T2  C1 . In other words, the inclined
crack has made the tensile force at point C a function for the moment at section A-B-D-E. This
shift in the tensile force must be considered in detailing the bar cut off points and in anchoring
the bars.

3.1.2.2. Behavior of beams with web reinforcement


Inclined cracking causes the shear strength of beams to drop below the flexural capacity. The
purpose of web reinforcement is to ensure that the full flexural capacity can be developed.

CHAPTER 3 – Limit State Design for Shear Page 64


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Prior to inclined cracking, the strain in the web reinforcement is equal to the corresponding strain
of the concrete. Because concrete cracks at a very small strain, the stress in the web
reinforcements prior to inclined cracking will not exceed 3 to 6 ksi. Thus, web reinforcements do
not prevent inclined cracks from forming; they come into play after the cracks have formed.

The forces in a beam with web reinforcements and an inclined crack are shown in Figure 3-4.

Figure 3-4 – Internal forces in a cracked beam with web reinforcements

The shear transferred by tension in the web reinforcements, v s , does not disappear when the
crack opens wider, so there will always be a compression force C1 and a shear force Vcy acting
on the part of the beam below the crack. As a result, T2 will be less than T1 , the difference
depending on the amount of web reinforcement. The force T2 will however, be larger than the
flexural tension T  M jd based on the moment at C

The loading history of such a beam is shown qualitatively in Figure 3-5.

CHAPTER 3 – Limit State Design for Shear Page 65


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 3-5 – Distribution of Internal shears in a beam with web reinforcement

The components of the internal shear resistance must equal the applied shear, indicated by the
upper 450 line. Prior to flexural cracking, the entire shear is carried by the uncracked concrete.
Between flexural and inclined cracking, the external shear is resisted by v cy , v ay , and v d .
Eventually, the web reinforcements crossing the crack yield, and v s stays constant for higher
applied shears. Once the web reinforcements yield, the inclined crack opens more rapidly. As the
inclined crack widens, v ay decreases further, forcing v d and v cy to increase at an accelerated
rate, until either a splitting (dowel) failure occurs, the compression zone crushes due to combined
shear and compression, or the web crushes.

Each of the components of this process except v s has a brittle load-deflection response. As a
result, it is difficult to quantify the contributions of v cy , v d , and v ay . In design, these are lumped
together as v c , referred to somewhat incorrectly as “the shear carried by the concrete.” Thus, the
nominal shear strength, v n , is assumed to be

Vn  Vc  Vs (22)

CHAPTER 3 – Limit State Design for Shear Page 66


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

3.1.3. FACTORS AFFECTING THE SHEAR STRENGTH OF BEAMS WITHOUT WEB


REINFORCEMENT
Beams without web reinforcement will fail when inclined cracking occurs or shortly afterwards.
For this reason, the shear capacity of such members is taken equal to the inclined cracking shear.
The inclined cracking load of a beam is affected by five principal variables, some included in
design equations and others not.

 Tensile strength of concrete

The inclined cracking load is a function of the tensile strength of the concrete, fct . The stress
state in the web of the beam involves biaxial principal tension and compression stresses. A
similar biaxial state of stress exists in a split-cylinder tension test, and the inclined cracking load
is frequently related to the strength from such a test. As discussed earlier, the flexural cracking
that precedes the inclined cracking disrupts the elastic-stress field to such an extent that inclined
cracking occurs at a principal tensile stress roughly half of fct for the uncracked section.

 Longitudinal reinforcement ratio, w

When the steel ratio, w , is small, flexural cracks extend higher into the beam and open wider
than would be the case for large values of w . And increase in crack width causes a decrease in
the maximum values of the components of shear, v d and v ay , that are transferred across the
inclined cracks by dowel action or by shear stresses on the crack surfaces. Eventually, the
resistance along the crack drops below that required to resist the loads, and the beam fails
suddenly in shear.

 Shear span to depth ratio, a/d

The shear span to depth ratio ,a/d or MV/d, affects the inclined cracking shears and ultimate
shears of portions of members with a/d less than 2.

 Lightweight aggregate concrete

Lightweight aggregate concrete has a lower tensile strength than normal weight concrete for a
given concrete compressive strength. Because the shear strength of a concrete member without
shear reinforcement is directly related to the tensile strength of a concrete, equations for shear
capacity must be modified for members constructed with light weight concrete.

 Size of beam

An increase in the overall depth of a beam with very little (or no) web reinforcement results in a
decrease in the shear at failure for a given fc , w , and a/d. The width of an inclined crack
depends on the product of the strain in the reinforcement crossing the crack and the spacing of

CHAPTER 3 – Limit State Design for Shear Page 67


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

the cracks. With increasing beam depth, the crack spacings and the crack widths tend to increase.
This leads to a reduction in the maximum shear stress that can be transferred across the crack by
aggregate interlock. An unstable situation develops when the shear stresses transferred across the
crack exceeds the shear strength. When this occurs, the faces of the crack slip, one relative to the
other.

 Axial forces

Axial tensile forces tend to decrease the inclined cracking load, while axial compressive forces
tend to increase it. As the axial compressive force is increased, the onset of flexural cracking is
delayed, and the flexural cracks do not penetrate as far into the beam. Axial tension forces
directly increase the tension stress, and hence the strain, in the longitudinal reinforcement. This
causes an increase in the inclined crack width, which, in turn, results in a decrease in the
maximum shear tension stress that can be transmitted across the crack. This reduces the shear
failure load.

A similar increase is observed in prestressed concrete beams. The compression due to


prestressing reduces the longitudinal strain, leading to a higher failure load.

 Coarse aggregate size

As the size (diameter) of the coarse aggregate increases, the roughness of the crack surfaces
increases, allowing higher shear stresses to be transferred across the cracks. In high strength
concrete beams and some light weight concrete beams, the cracks penetrate pieces of the
aggregate rather than going around them, resulting in a smoother crack surface. This decrease in
the shear transferred by aggregate interlock along the cracks reduces v c .

3.2. DESIGN OF BEAMS FOR VERTICAL SHEAR ACCORDING TO EN 1992-1-1-2004


For the verification of the shear resistance the following symbols are defined:

VRd ,c is the design shear resistance of the member without shear reinforcement
VRd ,s is the design value of the shear force which can be sustained by the yielding shear
reinforcement
VRd ,max is the design value of the maximum shear force which can be sustained by the
member, limited by crushing of the compression struts.
The shear resistance of a member with shear reinforcement is equal to:

VRd  VRd ,s (23)


In regions of the member where VEd  VRd ,c , no calculated shear reinforcement is necessary. VEd
is the design shear force in the section considered resulting from the external loading.

CHAPTER 3 – Limit State Design for Shear Page 68


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

In regions where VEd  VRd ,c , sufficient shear reinforcement should be provided in order that
VEd  VRd

The design shear force should not exceed the permitted maximum value VRd ,max , anywhere in the
member.

For members subject to predominantly uniformly distributed loading the design shear force need
not be checked at a distance less that d from the face of the support. Any shear reinforcement
required should continue to the support. In addition it should be verified that the shear at the
support does not exceed VRd ,max

3.2.1. MEMBERS NOT REQUIRING DESIGN SHEAR REINFORCEMENT


The design value for the shear resistance VRd ,c is given by:

VRd ,c  CRd ,c k 100 1fck   k1 cp  bw d (24)


13

 
with a minimum of

VRd ,c  vmin  k1 cp  bw d (25)


where:

fck is in MPa

200 (26)
k  1  2.0 with d in mm
d
Asl (27)
1   0.02
bw d

Asl is the area of the tensile reinforcement, which extends   l bd  d  beyond the
section considered
bw is the smallest width of the cross-section in the tensile area (mm)
 cp  NEd Ac  0.2fcd [MPa]
NEd is the axial force in the cross-section due to loading or prestressing in newtons (
NEd  0 for compression). The influence of imposed deformations on NE may be
ignored.
Ac Is the area of concrete cross section [mm2]
VRd ,c is in newtons

CHAPTER 3 – Limit State Design for Shear Page 69


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

The values of CRd ,c ,  min and k1 for use in a country may be found in its National Annex. The
recommended value for CRd ,c is 0.18  c , that for v min is 0.035k 3 2  fck1 2 and that for k1 is 0.15.

The design of members with shear reinforcement is based on a truss model. The angle θ should
be limited. The limiting values of cot  for use in a country may be found in its National Annex.
The recommended limits are 1  cot   2.5

3.2.2. MEMBERS REQUIRING DESIGN SHEAR REINFORCEMENT


For members with vertical shear reinforcement, the shear resistance, v Rd is the smaller value of :

Asw (28)
VRd ,s  zfywd cot 
s
and

VRd ,max  c bw z fcd  cot   tan  (29)


Where

Asw is the cross-sectional area of the shear reinforcement


s is the spacing of the stirrups
fywd is the design yield strength of the shear reinforcement
 follows from the expression below
For reinforced and prestressed members, if the design stress of the shear reinforcement is below
80% of the characteristic yield stress fyk ,  may be taken as:

  0.6 for fck  60 MPa (30)


  0.9  fck 200  0.5 for fck  60 MPa (31)
The value of tan c for use in a Country may be found in its National Annex. The recommended
value is 1 for non-prestressed structures.

The maximum effective cross-sectional area of the shear reinforcement Asw ,max is given by:

Asw ,max fywd


1 (32)
  c fcd
bw s 2
For members with inclined shear reinforcement, the shear resistance is the smaller value of

Asw (33)
VRd ,s  zfywd  cot   cot   sin
s
And


VRd ,max  c bw z fcd  cot   tan  1 cot 2   (34)

CHAPTER 3 – Limit State Design for Shear Page 70


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

The maximum effective shear reinforcement Asw ,max follows from:

1 (35)
Asw ,max fywd  c fcd sin
 2
bw s 1  cos 

3.2.3. ADDITIONAL TENSILE FORCE IN LONGITUDINAL REINFORCEMENT


The longitudinal tension reinforcement should be able to resist the additional tensile force caused
by shear.

The additional tensile force, Ftd , in the longitudinal reinforcement due to shear VEd may be
calculated from:

Ftd  0.5VEd  cot   cot   (36)


MEd z   Ftd should be taken not greater than MEd ,max z

3.2.4. MINIMUM AREA AND MAXIMUM SPACING OF SHEAR REINFORCEMENT


The ratio of shear reinforcement is given by

w  Asw /  s  bw  sin  (37)


where:

w is the shear reinforcement ratio


w should not be less than w ,min
Asw is the area of shear reinforcement within length s
s is the spacing of the shear reinforcement measured along the longitudinal axis of
the member
bw is the breadth of the web of the member
 Is the angle between shear reinforcement and the longitudinal axis
When, on the basis of the design shear calculation, no shear reinforcement is required, minimum
shear reinforcement should nevertheless be provided. The minimum shear reinforcement may be
omitted in members such as slabs (solid, ribbed or hollow core slabs) where transverse
redistribution of loads is possible. Minimum reinforcement may also be omitted in members of
minor importance which do not contribute significantly to the overall resistance and stability of
the structure.

The value of w ,min for beams for use in a Country may be found in its National Annex. The


recommended value is w ,min  0.08 fck f
yk

The maximum longitudinal spacing between shear assemblies should not exceed sl ,max .

CHAPTER 3 – Limit State Design for Shear Page 71


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

The value of sl ,max for use in a country may be found in its National Annex. The recommended
value is smax  0.75d 1 cot  

where  is the inclination of the shear reinforcement to the longitudinal axis of the beam.

3.2.5. PROCEDURE FOR DESIGN

 Step 1: Determine maximum applied shear force at support, VEd


V
 Step 2: Determine Rd ,max with cot   2.5
If VRd ,max  VEd cot   2.5 , go to step 6 and calculate required shear
 Step 3: reinforcement
If VRd ,max  VEd calculate required strut angle:
 Step 4:
 
  0.5sin1   Ed  0.20fck 1  fck 250   

 Step 5: If cot  is less than 1, re-size element, otherwise


Calculate amount of shear reinforcement required
 Step 6: Asw s   Ed bw  fywd cot   VEd  0.78dfyk cot  

 Step 7: Check min shear reinforcement and maximum spacing

CHAPTER 3 – Limit State Design for Shear Page 72


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

CHAPTER 4. SERVICEBILITY LIMIT STATE

4.1. INTRODUCTION
In chapter one of this course, the concept of serviceability limit state (SLS) has been introduced
briefly as part of the limit state approach. In this chapter a more elaborated and detailed
discussion is presented in accordance to EC2 with theoretical background.

In design of structures, the chief items of behavior of which are of practical significance are

1. The strength of the structure. i.e., the magnitude of loads which will cause the
structure to fail and

2. The deformations, such as deflections and extent of cracking, which the structure
will undergo when loaded under service conditions.

In the previous chapters we mainly dealt with the strength design of RC horizontal flexural
member. It is also important that member performance in normal service be satisfactory, when
loads are those actually expected to act, (i.e. when load factor is ≤1.0), which is not guaranteed
simply by providing adequate strength. Serviceability studies are carried out based on elastic
theory, with stress in both concrete and steel assumed to be proportional to strain. The concrete
on the tension side of the neutral axis may be assumed un-cracked, partially cracked, or fully
cracked, depending on the loads and material strengths.

The concept of Serviceability limit states has been introduced in chapter 1 and for RC structures
these states are often satisfied by observing empirical rules which affect the detailing and
dimensioning only. In some circumstances, however, it may be desired to estimate the behavior
of a member under working conditions, and mathematical methods of estimating deformations
and cracking must be used.

The major SLS for reinforced concrete structures are:

 stress limitation,
 excessive crack widths,
 excessive deflections, and
 undesirable vibrations.

Historically, deflections and crack widths have not been a problem for RC building structures.
With the advent high strength steel (fyk ≥ 400 MPa), the reinforcement stresses at service loads
have increased by about 50%. Since crack widths, deflections and fatigue are all related to steel
stress, each of these has become more critical.

CHAPTER 4 – Serviceability Limit State Page 73


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

4.2. ELASTIC ANALYSIS OF BEAM SECTIONS

4.2.1. SECTION UN-CRACKED


As long as the tensile stress in the concrete is smaller than the tensile strength of concrete (f ctk)
the strain and stress is the same as in an elastic, homogeneous beam. The only difference is the
presence of another material, i.e. the steel reinforcement. As it can be shown, in the elastic range,
for any given value of strain, the stress in the steel is 'n' times that of the concrete, where n
=Es/Ec is the modular ratio. In calculation the actual steel and concrete cross-section could be
replaced by a fictitious section (transformed section) thought of as consisting of concrete only. In
this section the actual steel area is replaced with an equivalent concrete area (nAs) located at the
level of the steel. Once the transformed section has been obtained, the beam is analyzed like an
elastic homogeneous beam.

A3

Compressive reinforcemnt

A1

Tensile reinforcemnt

A2

Figure 4-1 - Transformed Un-Cracked Section

4.2.2. SECTION CRACKED


When the tension stresses fct exceeds fctk, cracks form in the tension zone of the section. If the
concrete compressive stress is smaller than approximately 0.5fck and the steel has not reached the
yield strength, both materials continue to behave elastically.

At this stage, it is assumed that tension cracks have progressed all the way to the neutral axis and
that sections that are plane before bending remain plane in the bent member. This situation of the
section, strain and stress distribution is shown in the Error! Reference source not found.
below.

CHAPTER 4 – Serviceability Limit State Page 74


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

A3
Compressive reinforcemnt

A1

Tensile reinforcemnt

A2

Figure 4-2 - Transformed Cracked Section

4.3. STRESS CONTROL


The compressive stress in the concrete shall be limited in order to avoid longitudinal cracks,
micro-cracks or high levels of creep, where they could result in unacceptable effects on the
function of the structure.

Longitudinal cracks may occur if the stress level under the characteristic combination of loads
exceeds a critical value. Such cracking may lead to a reduction of durability. In the absence of
other measures, such as an increase in the cover to reinforcement in the compressive zone or
confinement by transverse reinforcement, it may be appropriate to limit the compressive stress to
a value k1fck in areas exposed to environments of exposure classes XD, XF and XS (refer chapter
2). If the stress in the concrete under the quasi-permanent loads is less than k2fck, linear creep can
be assumed. If the stress in concrete exceeds k2fck, non-linear creep should be considered (see
3.1.4)

Tensile stresses in the reinforcement shall be limited in order to avoid inelastic strain,
unacceptable cracking or deformation. Unacceptable cracking or deformation may be assumed to
be avoided if, under the characteristic combination of loads, the tensile stress in the
reinforcement does not exceed k3fyk. Where the stress is caused by imposed deformations, the
tensile stress should not exceed k4fyk. The mean value of the stress in prestressing tendons should
not exceed k5fpk.

Note: The recommended values k1, k2, k3, k4 and k5, which exempt to national annex

K1 0.6
K2 0.45
K3 0.8
K4 1.0
K5 0.75

CHAPTER 4 – Serviceability Limit State Page 75


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

4.4. SERVICEBILITY LIMIT STATE OF CRACKING

4.4.1. GENERAL INTRODUCTION


Serviceability limit sate refers to the performance of structures under normal service loads and is
concerned with the uses and/or occupancy of structures. The occurrence of cracks in reinforced
concrete is inevitable because of the low tensile strength of concrete. Structures designed with
low steel stresses at service load serve their intended function with very limited cracking. Crack
widths are of concern for three main reasons: aesthetic appearance, leakage and corrosion.

Aesthetic appearance:- The limits on aesthetic acceptability are difficult to set because
of the variability of personal opinion. The maximum crack width that will neither impair
a structure’s appearance nor create public alarm is probably in the range of 0.25 to 0.38
mm.
Leakage:- Crack control is important in the design of liquid-retaining structures. Leakage
is basically a function of the crack width.
Corrosion:- Concrete made from portland cement usually provide good protection for
reinforcement steel due its high alkalinity. Corrosion of the reinforcement happens when
an electrolytic cell is formed due to the carbonization of the concrete or chlorides
penetrate through the concrete reaches the bar surface. The time taken for this to occur
will depend on whether or not the concrete is cracked, the environment, the thickness of
the cover, and the permeability of the concrete. If the concrete is cracked, the time
required for a corrosion cell to be established is the function of the crack width.

It should be clearly understood that there are many causes of cracking and that only certain of
these lead to cracks that will be controlled by the provisions of chapter 7.3 of EC2. Chapter 7.3 is
concerned with cracks that form in hardened concrete either from restrained imposed
deformations, such as shrinkage or early thermal movements, or from the effects of loads.

The fundamental principle behind the provisions of the code is as follows. Crack control is only
possible where spread cracking can occur (i.e. the tensile strain is accommodated in multiple
cracks, or a crack accommodates only tensile strains that arise near the crack). For this to occur
there must be sufficient reinforcement in the section to ensure that the reinforcement does not
yield on first cracking. The rules for minimum reinforcement areas in 7.3.2 are aimed at ensuring
that this requirement is met. Provided this minimum is present, crack widths can normally be
controlled by simple detailing rules.

4.4.2. CAUSES OF CRACK


I. Load induced cracks: Tensile stress induce by loads, moments, and shear cause
distinctive crack patterns as shown below:

CHAPTER 4 – Serviceability Limit State Page 76


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 4-3 - Load induced cracks

II. Heat of hydration cracking: a frequent cause of cracking in structures is restrained


contraction resulting from the cooling down to ambient temperatures of very young
members which have expanded due to heat of hydration which developed as the concrete
was setting. A typical heat of hydration cracking pattern of a wall cast on foundation
concrete is shown below. Such cracking can be controlled by controlling the heat rise due
to the heat of hydration and the rate of cooling, or both; by placing the wall in short
lengths; or by reinforcements considerably in excess of normal shrinkage reinforcement.

CHAPTER 4 – Serviceability Limit State Page 77


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Figure 4-4 - Load induced cracks

III. Plastic slumping cracks: plastic shrinkage and slumping of the concrete occurs as newly
placed concrete bleeds and surface dries, results in settlement cracks along the
reinforcement as shown in Figure 5 (a) , or a random cracking pattern, referred to as map
cracking shown in Figure 5 (b). These types of cracks can be avoided by proper mix
design and by preventing rapid drying of the surface during the first hour or so after
placing. Map cracking can also occur due to alkali-aggregate reaction.

IV. Cracks caused by corrosion: rust occupies two to three times the volume of the metal
from which it is formed. As a result, if rusting occurs, a bursting force is generated at the
bar location which leads to splitting cracks and eventual loss of cover. Such cracking
looks similar to bond cracking (Figure (c)) and may accompany bond cracking.

Figure 4-5 - Load induced cracks

4.4.3. CRACK CONTROL


Crack control in accordance to EC2 may be achieved by one of the following approaches:

I. Limiting the maximum bar diameter using Table 7.2

CHAPTER 4 – Serviceability Limit State Page 78


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

II. Limiting the maximum bar spacing using Table 7.3 (this table is not applicable for
restraint loading)
III. Calculating cracks to ensure they are within limits (Table 7.1)

For Notes see Table 7.2N

CHAPTER 4 – Serviceability Limit State Page 79


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

4.4.3.1. Minimum area of Reinforcement


For most purposes, thermal and shrinkage cracking can be controlled within acceptable limits by
the use of minimum reinforcement quantities. The provision of the minimum steel area ensures
that the reinforcement does not yield when the concrete in the tension zone cracks with a sudden
transfer of stress to the reinforcement. This could cause the uncontrolled development of few
wide cracks. Whenever this minimum area is provided, then yield should occur and cracking will
then be distributed throughout the section with greater number of cracks but of lesser width.
As,min is given by the expression.

As ,min  kc kfct , eff Act / fyk Equation i


Where
As ,min is the minimum area of reinforcing steel within the tensile zone
A ct is the area of concrete within tensile zone. The tensile zone is that part of the
section which is calculated to be in tension just before formation of the first crack
s is the absolute value of the maximum stress permitted in the reinforcement
immediately after formation of the crack. This may be taken as the yield strength
of the reinforcement, fyk . A lower value may, however, be needed to satisfy the
crack width limits according to the maximum bar size (Table 7.2) or the maximum
bar spacing (Table 7.3)
fct , eff is the mean value of the tensile strength of the concrete effective at the time
when the cracks may first be expected to occur:
fct,eff = fctm or lower, (fctm (t)), if cracking is expected earlier than 28 days
k is the coefficient which allows for the effect of non-uniform self-equilibrating
stresses, which lead to a reduction of restraint forces
= 1,0 for webs with h  300 mm or flanges with widths less than 300 mm
= 0,65 for webs with h  800 mm or flanges with widths greater than 800 mm
intermediate values may be interpolated
kc is a coefficient which takes account of the nature of the stress distribution within
the section immediately prior to cracking and of the change of the lever arm:

CHAPTER 4 – Serviceability Limit State Page 80


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

For pure tension


k c  1.0
For bending or bending combined with axial forces:
-For rectangular sections and webs of box sections and T-sections:
 c 
k c  0.4 1  1
 k1  h / h  fct , eff 
*

-For flanges of box sections and T-sections:


Fcr
k c  0.9  0.5
Act fct , eff
Where
 c is the mean stress of the concrete acting on the part of the section under
consideration:
NEd
c 
bh
NEd is the axial force at the serviceability limit state acting on the part of the
cross-section under consideration (compressive force positive). NEd should
be determined considering the characteristic values of prestress and axial
forces under the relevant combination of actions
h* h*  h for h < 1.0 m
h*  1.0m for h  1.0 m
k1 is the coefficient considering the effects of axial forces on the stress distribution:
k1= 1.50 if NEd is a compressive force
2h*
k1= if NEd is a tensile force
3h
-For flanges of box sections and T-sections:
Fcr
k c  0.9  0.5
Act fct , eff
Where
Fcr is the absolute value of the tensile force within the flange immediately prior
to cracking due to the cracking moment calculated with fct,eff .

4.4.3.2. Control of Cracking without Direct Calculation


As discussed earlier crack control can be achieved by either limiting the maximum bar diameter
using Table 7.2 or limiting the maximum bar spacing using Table 7.3. But the maximum bar
diameter (Table 7.2) may be modified as follows:

CHAPTER 4 – Serviceability Limit State Page 81


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

kc hcr
s   * s  fct , eff / 2.9 Bending (at least part of section in compression) Equation ii
2h  d 

kc hcr
s   * s  fct , eff / 2.9 Tension (all of section iunder tension) Equation iii
h  d 
Where
s is the adjusted maximum bar diameter
 *s is the maximum bar size given in the Table 7.2
h is the overall depth of the section
hcr is the depth of the tensile zone immediately prior to cracking, considering the
characteristic values of prestress and axial forces under the quasi-permanent
combination of actions
d is the effective depth to the centroid of the outer layer of reinforcement

4.4.4. CRACK WIDTH CALCULATION


According to EC2 7.3.4 the following procedure shall be employed to compute the characteristic
crack width, wk, may be obtained from the relation:

wk  sr ,max  sm   cm  Equation iv

Where
s r ,max is the maximum crack spacing
 cm is the mean strain in the reinforcement under the relevant combination of loads,
including the effect of imposed deformations and taking into account the effects of
tension stiffening. Only the additional tensile strain beyond zero strain in the
concrete is considered
 sm is the mean strain in the concrete between cracks

fct , eff
 s  kt
  , eff
1  e  ,eff  s
 sm   cm   0.6 Equation v
Es Es

CHAPTER 4 – Serviceability Limit State Page 82


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Where
s is the stress in the tension reinforcement assuming a cracked section.
e is the ratio ratio Es /Ecm
  ,eff is the effective reinforcement ratio.
As
  ,eff =
Ac , eff
As is area of tension reinforcement
Ac , eff is the effective tension area. A c,eff is the area of concrete surrounding the tension
reinforcement of depth, hc,ef , where hc,ef is the lesser of 2.5(h-d), (h-x)/3 or h/2
(see Figure 7.1)

sr ,max  3.4c  0.425k1k2 /  , eff (Equation 7.11)Equation vi

CHAPTER 4 – Serviceability Limit State Page 83


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

Where
c is the cover to the reinforcement
k1 is a coefficient which takes account of the bond properties of the bonded
reinforcement:
= 0.8 for high bond bars
= 1.6 for plain bars
k2 is a coefficient which takes account of the distribution of strain:
= 0.5 for bending
= 1.0 for pure tension
 Bar diameter

When spacing > 5  c   / 2


(Equation 7.14)Equation vii
sr ,max  1.3  h  x 

4.5. SERVICEBILITY LIMIT STATE OF DEFLECTION

4.5.1. GENERAL INTRODUCTION


In addition to limitation on cracking, described in the preceding sections, it is usually necessary
to impose certain controls on deflections of beams to ensure serviceability. Excessive deflections
can lead to cracking of supported walls and partitions, ill-fitting doors and windows, poor roof
drainage, misalignment of sensitive machinery and equipment, or visually offensive sag. It is

CHAPTER 4 – Serviceability Limit State Page 84


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

important, therefore, to maintain control of deflections, in one way or another, so that members
designed mainly for strength at prescribed overloads will also perform well in normal service.

According to EC2, the deflection of a structure or any part of the structure shall not adversely
affect the proper functioning or appearance of the structure. This may be ensured either by
comparing a calculated deflection, according to 7.4.3, with a limit value or by limiting the
span/depth ratio, according to 7.4.2.

4.5.2. CASES WHERE CACLCULATION MAY BE OMITTED


Generally, it is not necessary to calculate the deflections explicitly as simple rules, for example
limits to span/depth ratio may be formulated, which will be adequate for avoiding deflection
problems in normal circumstances. More rigorous checks are necessary for members which lie
outside such limits, or where deflection limits other than those implicit in simplified methods are
appropriate.

1  o  o  2 
3

 K 11 1.5 fck  3.2 fck   1  if   o Equation viii


d     

1  o 1 
 K 11 1.5 fck  fck  if    o Equation ix
d     ' 12 '

Where
l is the limit span/depth
d
K is the factor to take into account the different structural systems
0 is the refrence reinforcement ratio = fck 10-3
 is the required tension reinforcement ratio at mid-span to resist the moment due to the design loads
(at support for cantilevers)
 ' is the required compression reinforcement ratio at mid-span to resist the moment due to the design loads
(at support for cantilevers)
fck is in MPa units

Equations i and ii are to be modified as follows:

a) the values obtained should be multiplied by 310 /  s , which will normally be


conservative to assume that:

CHAPTER 4 – Serviceability Limit State Page 85


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

310 /  s  500 / (fyk As , req / As , prov ) Equation x

Where
 s is the tensile steel stress aat mid-span (at support for cantilevers) under the design service load
A s,prov is the area of steel provided at this section
A s,req is the area of steel required at this section for ultimate limit state

b) For flanged sections where the ratio of the flange breadth to the rib breadth exceeds 3,
the values of l/d given by equation I should be multiplied by 0.8.
c) For beams and slabs, other than flat slabs, with spans exceeding 7m, which support
partitions liable to be damaged by excessive deflections, the values of l/d given by
Equation I should be multiplied by 7/ leff (leff in metres)
d) For flat slabs where the greater span exceeds 8,5 m, and which support partitions liable
to be damaged by excessive deflections, the values of l/d given by Expression (7.16)
should be multiplied by 8,5 / leff (leff in metres). (NOT APPLICABLE FOR THIS
COURSE!)

Table 3Basic ratios of span/effective depth for reinforced concrete members without axial
compression

Concrete highly Concrete lightly


Structural System K
stressed ρ=1.5% stressed ρ=0.5%
Simply supported beam, one- or 1.0 14 20
two-way spanning simply
supported slab
End span of continuous beam or 1.3 18 26
one-way continuous slab or two-
way spanning slab continuous over
one long side
Interior span of beam or one-way 1.5 20 30
or two-way spanning slab
Slab supported on columns without 1.2 17 24
beams (flat slab) (based on longer
span)
Contilever 0.4 6 8
Note 1: The values given have been chosen to be generally conservative and calculation may
frequently show that thinner members are possible.
Note 2: For 2-way spanning slabs, the check should be carried out on the basis of the shorter span.
For flat slabs the longer span should be taken.
Note 3: The limits given for flat slabs correspond to a less severe limitation than a mid-span
deflection of span/250 relative to the columns. Experience has shown this to be satisfactory. ).
(NOT APPLICABLE FOR THIS COURSE!)

CHAPTER 4 – Serviceability Limit State Page 86


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

4.5.3. DEFLECTION LIMITS


Deflections are limited for the following reasons:

i. Excessive deflections are unsightly and alarming. EC2 restricts total deflection to
span/250.
ii. To avoid damage to cladding, partitions and finishes due to increments in deflection
following their construction. EC2 limits deflections after construction of finishes to
span/500.
iii. Both construction tolerances and deflections need to be considered in the design of
fixings for cladding systems and partitions. In practice it can be difficult to separate
construction tolerances from deflections.

4.5.4. CALCULATION FOR DEFLECTION


The requirements for the calculation of deflections are given in Section 7.4.3 and Appendix C of
EC2.

Two limiting conditions are assumed to exist for the deformation of concrete sections

1) Uncracked

2) Cracked.
Members which are not expected to be loaded above the level which would cause the tensile
strength of the concrete to be exceeded, anywhere in the member, will be considered to be
uncracked. Members which are expected to crack will behave in a manner intermediate between
the uncracked and fully cracked conditions.

For members subjected dominantly to flexure, the Code gives a general equation for obtaining
the intermediate value of any parameter between the limiting conditions

  (1  )I  II Equation xi

Where
 is the parameter being considered
I and II are the values of the parameter calculated for the uncracked and fully
cracked conditions respectively
 is a distribution coefficient ( allwing for tension sttifening at a section ) given by

CHAPTER 4 – Serviceability Limit State Page 87


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

2
 
  1   sr  Equation xii
s 

  0 for uncracked sections


 is a coefficent taking account of the influence of the duration of the loading
or of the repeated loading on the average strain
=1.0 for a single short-term loading
= 0.5 for sustained loads or many cycles of repeated loading
s is the stress in the tension reinforcemnt calculated on the basisi of cracked section
 sr is the stress in the tension reinforcemnt calculated on the basis of cracked section
under the loading conditions causing first craking

Note: σsr/ σs may be replaced by Mcr/M for flexure or Ncr/N for pure tension, where
Mcr is the cracking moment and Ncr is the cracking force.

The effects of creep are catered for by the use of an effective modulus of elasticity for
the concrete given by
 Ecm 
Ec,eff =   Equation xiii
1  (, to ) 

Where:
 (, to ) is the creep coefficent relevant for the load and time interval

Curvatures due to shrinkage may be assessed from


1 S
=  cs e Equation xiv
rcs I
Where:
1/ rcs is the curvature due to shrinkage
 cs is the free shrinkage strain
S is the first moment of area of the reinforcemnt about the centroid of the section
I second moment of area of the section
e is the effective modular ratio

Shrinkage curvatures should be calculated for the uncracked and fully cracked conditions and the
final curvature assessed by use of Equation iv.

CHAPTER 4 – Serviceability Limit State Page 88


AAiT, School of Civil and Environmental Engineering Reinforced Concrete I

In accordance with the Code, the rigorous method of assessing deflections is to calculate the
curvatures at frequent sections along the member and calculate the deflections by numerical
integration.

The simplified approach, suggested by the Code, is to calculate the deflection assuming firstly
the whole member to be uncracked and secondly the whole member to be cracked. Equation iv is
used to assess the final deflection,

CHAPTER 4 – Serviceability Limit State Page 89

You might also like