Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

RC-I Design of Beams for Shear

CHAPTER 4
DESIGN OF BEAMS FOR SHEAR

Introduction
Beams resist loads primarily by means of internal moments, M, and internal shears, V. In the
design of reinforced concrete members, flexure is usually considered first, (i.e. sections are
proportioned and areas of longitudinal reinforcement determined for the moment M), because
flexural failure is ductile. The beams are then designed for shear. Because shear failure is
frequently sudden and brittle, the design for shear must ensure that shear strength equals or
exceeds the flexural strength at all points in the beam.
4.1 Theoretical Background
Stresses in an Uncracked Elastic Beam
The role of shear stresses is easily visualized by the performance under load of the laminated
beam of Figure B.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 B.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 B.1b.
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 B.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 B.1d shows a differential length of a single-piece rectangular beam acted upon by a
shear force of magnitude V. Vertical equilibrium is provided by the vertical shear stresses, ν.
For a homogeneous, elastic, uncracked beam, the shear stress, ν, is given by;
VQ
 (B.1)
Ib
The shear stress is zero at the outer fibers and has a maximum of 1.5V/bh at the neutral axis,
the variation being parabolic.

Figure B.1: Shear in homogeneous elastic beams.

Unity University 1 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

If a stress element located at the neutral axis is isolated as shown in Figure B.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 are equal in magnitude to the vertical shear stresses.
Recall that beams are subjected to flexural normal stress, f, and shearing stress, ν. Therefore,
the principal stresses are;

 
2 f
 Major  x  2x   xy 2   2f  2
2 2
2

(B.2)
 Minor 
x
2
 
x 2 f
 2   xy   2  
2

2
f 2
 2

The stress element in Figure B.2b is in a state of pure shear (f = 0) and the principal stresses
are equal (σMaj = - σMin = ν) and are located at an element cut at 45° as shown in Figure B.2c.

Figure B.2: stresses and stress trajectories in a homogenous uncracked beam.


If an element is located neither at the neutral axis nor at the outer edges (element-2), its vertical
faces are subject not only to the shear stresses but also to bending stresses and it is said to be
in a biaxial state as shown in Figure B.2d. The principal stresses are shown in Figure B.2e.
Since the magnitudes of the shear stresses and the bending stresses 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 B.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 45° to the axis. In the vicinity of the
outer fibers they are horizontal near midspan.

Unity University 2 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

Planes of principal tensile stress (these are the compression trajectories) are steep near the
bottom of the beam and flatter near the top.
In concrete beams with longitudinal flexural reinforcement, but no shear reinforcement, two
types of cracks can be seen. The vertical cracks occurred first, due to flexural stresses. These
start at the bottom of the beam where the flexural stresses are the largest. The inclined cracks
near the ends of the beam are due to combined shear and flexure. These are commonly referred
to as inclined cracks, shear cracks, or diagonal tension cracks. Such a crack must exist before
a beam can fail in shear.
Although there is similarity between the planes of maximum principal tensile stress and the
cracking pattern, it is by no means perfect, because in RC beams flexural cracks generally occur
before the principal tensile stress at mid height become critical. Once the flexural cracks has
occurred, the tensile stress perpendicular to the cracks drops to zero. To maintain equilibrium,
a major redistribution is necessary. As a result, the onset of inclined cracking in a beam cannot
be predicted from the principal stresses unless shear cracks precedes flexural cracking. This
very rarely happens in RC, but it does occur in some pre-stressed beams.
Shear transfer of reinforced concrete beams heavily relies on the tensile and compressive
stresses of the concrete. Most of the time, the problem of concrete in shear design is not shear
stress exceeding the shear strength of the concrete, but the major tensile principal stress
exceeding the tensile strength of concrete due to the low tensile strength. When the tensile
stress exceeds the tensile strength of concrete, cracks will form, and the formation of cracks
results a complex pattern of stresses.
4.2 Mechanism of Shear Resistance in Concrete Beams without Shear Reinforcements
After formation of cracks the stress distribution is different from the ones discussed above.
Figure B.3a shows a cracked beam with longitudinal flexural reinforcement, but no shear
reinforcement.

Figure B.3: Calculation of average shear stress between cracks

Unity University 3 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

The concrete below the neutral axis in a cracked reinforced concrete beam is in a state of pure
shear because tensile flexural stress is zero. The equilibrium of the section of beam between
two cracks (Fig. B.3b) can be written as;
M M  M M
T and T  T   T  (B.3)
Z Z Z
From the moment equilibrium of the element;
Vx
M  Vx  T  (B.4)
Z
If the shaded portion of Fig B.3b is isolated, the force ΔT must be transferred by horizontal
shear stress on the top of the element. The average value of these stresses below the top of the
crack is;
T V
 or   (B.5)
bw x bw Z

Where bw is the width of the beam (width of the web in case of T-beam)
The distribution of the average horizontal shear stress is shown in Fig B.3d. Since the vertical
shear stresses on an element are equal to the horizontal shear stresses on the same element, the
vertical shear stress distribution will also be as shown in Fig B.3d.
4.2.1 Internal Forces in Concrete Beams without Shear Reinforcements
When a beam is subjected to flexure and shear, the shear resistance in the absence of shear
reinforcement is contributed by the concrete compression zone, mechanical interlock of
aggregate at the crack and dowel action of the longitudinal flexural reinforcement. Figure B.4
shows mechanism of shear resistance across an inclined crack in a beam without shear
reinforcement (stirrups).

Figure B.4: Internal forces in a cracked beam without stirrups.


Shear resistance along A-B-C is provided by the sum of; shear in the compression zone Vcz, the
vertical component of force due to aggregate interlock Vay, and force due to dowel action of
the longitudinal reinforcement Vd.
Immediately after inclined cracking occur, as much as 40 to 60% of the total shear is carried
by Vd and Vay together.

Unity University 4 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

Considering portion D-E-F of the beam below the crack and summing moments about the
reinforcement at point E shows that Va and Vd will cause a moment about E in clockwise
direction which should be balanced by the moment due to compression force C’1.
Horizontal force equilibrium on vertical face A-B-D-E shows that T1=C1+C’1 and finally T1
and C1+C’1 must equilibrate the external moment at this section.
As the crack widens, Va decreases, and much of the resistance is provided by Vcz and Vd. As Vd
gets larger it leads to splitting crack in the concrete along the reinforcement. When this crack
occurs Vd drops approaching to zero. When Va and Vd disappear, so do V’cz and C’1, with the
result that all the shear and compression are transmitted in the depth AB above the crack. This
may cause crushing of concrete in region AB.
It is important to note also that, if C’1 =0, then T2 = T1, and as a result, T2=C1. In other words,
the inclined crack has made the tensile force at point C to be a function of the moment on the
vertical section A-B-D-E. This shift in tensile force must be considered when determining bar
cutoff points and when anchoring bars.
For a typical RC beam, the approximate proportions of the forces are:-
Vcz = 20 – 40% Vd = 15 – 25% Vay = 35 – 50%
It has been found that the dowel action is generally the first to reach its capacity followed by
failure of the aggregate interlock, which is followed by shear failure of the concrete in
compression (abruptly & explosively). However, the precise proportion is difficult to establish
and the shear strength is represented by a single expression accounting for all mechanisms.
4.2.2 Factors Affecting the Shear Strength of Beams Without Shear Reinforcement
Beams without Shear 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 several factors, Including;
 The Tensile Strength of Concrete,
 Flexural reinforcement ratio (ρ).
 Shear span av to effective depth d ratio (av/d)
 Depth of the member
 Type of Aggregate
 Axial Force in the member
Tensile strength of concrete: - Shear resistance increases with increasing tensile strength.
Increasing the concrete strength increases the aggregate interlock capacity and also the capacity
of the uncracked portion of the beam.
Flexural reinforcement ratio (ρ): - This affects the shear capacity by restraining the width of
the cracks and thus enhancing the shear carried by the aggregate interlock along the cracks. It
also naturally increases the shear capacity due to dowel action and increases the depth of the
section in compression. Therefore, shear resistance increases with increasing reinforcement
ratio, but the rate of increase reduces as reinforcement ratio increases. To model this behavior,
ES EN 2 employs a cube root relationship between shear resistance and steel percentage up to
a maximum value of ρ of 0.02.

Unity University 5 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

Shear span av to effective depth d ratio (av/d): - Shear span av is defined as the distance between
the support and the major concentrated load acting on the span. Note that av = M/V. It affects
the inclined cracking shears and ultimate shears of portions of members with av/d less than 2.
Depth of the member: - this have a significant influence on shear resistance over and above that
expected from normal geometrical scaling (i.e. there is a size effect). Tests indicate that deeper
beams have proportionally lower shear capacity compared to shallow beams. The reason for
this is not clear but it is thought it might have some thing to do with lower aggregate interlock
capacity. Most recent codes of practice therefore have a term in their equations to allow for this
and which gives a higher shear strength for shallow members.
Type of aggregate: - This affects the shear resisted by aggregate interlock. For example,
lightweight aggregate concrete has approximately 20% lower shear capacity compared to
normal weight concrete.
Axial Force: - 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.
Empirical relations are given in codes which may consider all of these factors or only some.

Unity University 6 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

4.3 Shear Resistance of Concrete Beams with Shear Reinforcement


In the previous section we saw that formation of diagonal cracks is followed by widening of
cracks and brittle compression failure. This type of failure can be suppressed and development
of full flexural capacity can be ensured using shear reinforcement.

Figure B.6: Inclined cracks and types of shear reinforcements (stirrups).


Shear reinforcement can be provided in the form of; vertical stirrups (Fig. B.6d), inclined
stirrups (Fig. B.6c) or bent up longitudinal bars. The most commonly used type is vertical
stirrup. The use of bent bars has almost disappeared. If inclined stirrups are used, they should
form an angle, α, between 45° and 90° to the longitudinal axis of the structural element. Inclined
stirrups cannot be used beams resisting shear reversal such as building resisting seismic loads.
Before the formation of inclined cracking, the strain in the stirrups is equal to the corresponding
strain of the concrete. Because concrete cracks at a very small strain, the stress in the stirrups
before formation of inclined cracks is very small. Thus, stirrups do not prevent inclined cracks
from forming; they come into play only after these cracks have formed. The forces in a beam
with stirrups and an inclined crack are shown in Fig. B.7.

Figure B.7: Internal force in a cracked beam with stirrups

Unity University 7 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

The forces in a beam with stirrups & an inclined crack can be obtained from a mathematical
model called “The Truss-Analogy Model”. In this model, the stirrups are modeled as vertical
tension members, the longitudinal flexure rebars as horizontal tension members, the concrete
diagonals between cracks as diagonals compression members (struts inclined at θ) & the
concrete in flexural compression as top horizontal compression members as shown in Fig. B.8.

Figure B.8: The Truss-Analogy Model.

Figure B.9: Forces in stirrups & compression diagonals.

Unity University 8 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

If we consider the free-body diagram cut by section A–A parallel to the diagonals in the
compression field region as shown in Fig. B.9a, the entire shear force is resisted by tension
forces in the stirrups crossing this section.
The horizontal projection of section A–A is Zcotθ, and the number of stirrups it cuts is (Zcotθ)/S.
Since the force in one stirrup is Aswfywd; the shear resisted by the stirrups is;
V  ( No. of stirrups) *( Force in each stirrup)

Zcotθ
V * Asw f ywd (B.6)
S
Where, S = center to center spacing between shear reinforcement.
Asw = the area of shear reinforcement = number of legs*area of single leg
fywd = the design yield stress of the shear reinforcement.
Now consider the free-body diagram cut by section B–B parallel to the stirrups as shown in
Fig. B.9b. Here, the vertical force, V, acting on the section has been replaced by diagonal
compression force FD and axial tension force Nv as shown in Fig. B.9c. The stress in the
diagonal struts, σc, is the force FD distributed over an inclined area of bw*Zcosθ. That is;
FD V V
c  since FD   c 
bw Zcosθ sinθ sinθ bw Zcosθ

V  1  V
 c    tanθ   c   cotθ +tanθ  this gives;
bw Z  tanθ  bw Z
 cbw Z
V (B.7)
 cotθ +tanθ 
This is the maximum shear force that can be resisted without crushing the concrete in the struts
As shown in Fig. B.9b, due to the shear force V, the top and bottom chords support a tensile
force of magnitude 0.5Nv=0.5Vcotθ. Therefore, the additional tensile force, ΔFt, in the
longitudinal reinforcement is;
Ft  0.5Vcotθ (B.8)

The design forms of Equations B.6, B.7 and B.8 appear in ES EN-2 as Equations (6.8), (6.9)
and (6.18) respectively. They are shown in the following sections.
In shear design, there are two points where methods in various building codes differ. The 1st is
in the choice of a value for the strut angle, θ. The 2nd is whether the shear rebar should carry
all the shear force or whether part of the shear can be considered to be carried by the concrete.
Internationally, by far the most common approach is to assume a fixed strut angle of θ=45° (i.e.
cotθ=1), and to assume the shear is carried by both the concrete & the shear rebar. It is found
that in this method, the shear capacity of a beam with shear reinforcement is underestimated.
ES EN-2 adopts a variable strut inclination method, in which, all the shear is assumed to be
carried by the shear rebar, but the strut angle, θ, can take any value between 21.8° ≤ θ ≤ 45°
(that is 1 ≤ cotθ ≤ 2.5). This variable strut angle approach is considered to be the more rigorous
of the two methods, and is also slightly more economical in some cases. However, as written

Unity University 9 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

in ES EN-2, it is open to a misunderstanding. The code implies that the designer may select
any strut angle between the specified limits. This concept of free choice does not, however,
reflect the behaviour of a beam. Beams will fail in a manner corresponding to a strut angle of
roughly 21.8° unless constrained by the detailing or the geometry of the system to fail at some
steeper angle. A steeper-angled failure could be induced either by the way in which the tension
steel was curtailed or where the load is so close to the support that only a steeper failure can
occur or where the shear strength is limited by the crushing strength of the strut.
4.4 Design of Beams for Shear as per ES EN-2
For the verification of the shear resistance the following symbols are defined;
VEd = the design shear force in the section resulting from external loading and prestressing.
VRd,c = the design shear resistance of the member without shear reinforcement.
VRd,s = the design value of the shear force which can be sustained by the yielding shear rebar.
VRd,max = the design value of the maximum shear force which can be sustained by the member,
limited by crushing of the compression struts.
The following rules apply;
1. The shear resistance of a member with shear reinforcement is equal to: VRd = VRd,s
2. In regions where VEd ≤ VRd,c, no calculated shear reinforcement is necessary.
3. 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 (e.g. lintels with span ≤ 2 m) which
do not contribute significantly to the overall resistance and stability of the structure.
4. In regions where VEd > VRd,c , sufficient shear reinforcement should be provided in
order that VEd ≤ VRd.
5. The design shear force should not exceed the permitted maximum value VRd,max,
anywhere in the member.
6. For members subjected 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.
7. The longitudinal tension reinforcement should be able to resist the additional tensile
force caused by shear.
Note that the critical section for shear is at a distance d from the face of support. Therefore, the
design shear force, VEd, is the shear force at a distance d from the face of support due to external
loading and prestressing.
The equations for shear resistance for members without or with shear rebar as per ES EN-2 are
presented below.

Unity University 10 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

Members Not Requiring Design Shear Reinforcement


Taking the factors that contribute to the shear resistance of concrete into account, ES EN-2
gives the following equation for the shear resistance of members without shear reinforcement:

C k 100 f  1/3


bwd
VRd ,c   Rd ,c 1 ck
------------------------------------ (4.1)
vmin bwd
This equation is applicable if there is no significant axial force.
Where,
VRd,c = design shear resistance of a member without shear reinforcement.
CRd,c = 0.18/c =0.18/1.5 = 0.12
bw = the smallest width of the cross section in the tension area.

vmin  0.035  k1.5  fck

 200  As
1  
k  d where d is in mm. 1   bw d
2 0.02
 
As = area of tensile reinforcement extending at least lbd + d beyond the section considered (see
Figure 4.1)

Figure 4.1: Definition of As.


Note that if the design shear force, VEd, is less than or equal to VRd,c, shear reinforcement is not
required theoretically.
If the section supports an axial load (as in columns, walls and prestressed members), the effect
of the axial load should be included and hence, Equation 4.1 will become;

C k 100 f 1/3 b d  0.15 b d  N Ed



  Rd ,c where  cp   Ac
1 ck w cp w
VRd ,c
vmin bwd  0.15 cpbwd 0.2 f
 cd

NEd = the axial force in the cross-section due to loading or prestressing [in N]. NEd is
+ve for compression. The influence of imposed deformations on NEd can be ignored.
Ac = the area of concrete cross section [mm2]

Unity University 11 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

Members Requiring Design Shear Reinforcement


For members with shear reinforcement, the shear resistance is governed by either by the
yielding of shear reinforcement or by the crushing of compression struts.
Shear resistance governed by yielding of shear reinforcement, VRds, is given by;
A
VRd ,s  sw zf ywd cot  ----------------------------------------- (4.2)
S
Shear resistance limited by the crushing of compression struts, VRd,max, is given by;
 b z fcd
VRd ,max  cw w --------------------------------------- (4.3)
 cot   tan 
Notice that these equations are the same as Eqn. B.6 and Eqn. B.7 (ES EN-2 limits the concrete
stress σc to;  c   cw f cd ).
Where,
Asw = cross sectional area of the shear reinforcement.
S = center to center spacing of the shear reinforcement.
z = lever arm = 0.9d.
fywd = yield strength of the shear reinforcement.
θ = angle of inclination of the strut. This is limited to 21.8° ≤ θ ≤ 45° (that is 1 ≤ cotθ ≤ 2.5).
αcw = 1 = a coefficient that takes into account the state of stress in the compression chord.
ν = strength reduction factor for concrete cracked in shear. The recommended value is;
 f 
  0.6 1  ck 
 250 
fcd = compressive strength of concrete. Note that for shear design it is given by = fck/γc.
If inclined shear reinforcement is used, equation 4.2 & 4.3 will become.
A
VRd ,s  sw zf ywd  cot   cot   sin  --------------------------- (4.4)
S
 b z fcd  cot   cot  
VRd ,max  cw w ---------------------------- (4.5)

1  cot 2  
Additional Tensile Force in The Longitudinal Reinforcement
The additional tensile force, ΔFtd, in the longitudinal reinforcement due to shear VEd can be
calculated from:
Ftd  0.5VEd cotθ
(MEd/z)+ΔFtd should be ≤ MEd,max/z, where MEd,max is the maximum moment along the beam.
If inclined shear reinforcement is used, the above eqnuation becomes;
Ftd  0.5VEd  cotθ - cot 

Unity University 12 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

Design of Shear Reinforcement


Shear reinforcement may consist of a combination of;
 Links enclosing the longitudinal tensile reinforcement and the compression zone.
 Shear assemblies of cages, ladders, etc. which do not enclose the longitudinal
reinforcement, but which are properly anchored in the compression and tension zones.

Figure 4.3: Shear stirrups and links.


The required area of shear reinforcement in a certain spacing, Asw/S, can be obtained from eqn.
4.2 by setting the shear resistance, VRds, to be equal to the design applied shear force, VEd, and
rearranging;
Asw VEd
 ----------------------------------------- (4.6)
S zf ywd cot 
Shear reinforcement is usually expressed as the spacing, S, between the stirrups. The spacing
can be obtained from the result of the above equation.
Minimum Area and Maximum Spacing
Minimum shear reinforcement should be provided even if shear reinforcement is not required
theoretically. The minimum shear reinforcement ratio is given by;
0.08 f ck
ρw,min  ----------------------------------------------- (4.7)
f yk
Since ρw,min = Asw/(S∙bw), the minimum shear reinforcement in a certain spacing will be;

 Asw  0.08bw fck


   ----------------------------------------- (4.8)
 S min f yk
The longitudinal spacing between shear reinforcement assemblies should not exceed Smax.
Smax  0.75d --------------------------------------------- (4.9)
The transverse spacing of the legs in a series of shear links should not exceed St,max.
0.75d
St ,max   ------------------------------------------- (4.10)
600mm

Unity University 13 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

Procedure for Design


1) Calculate the design applied shear force, VEd.
2) Calculate VRd,c.
3) If VEd ≤ VRd,c, then shear reinforcement is not required. Hence go to step-5.
4) If VEd > VRd,c, shear reinforcement is required and it should be calculated.
4.1) Calculate VRd,max for θ = 45° (i.e. cotθ=1) and for θ = 21.8° (i.e. cotθ=2.5). These are;
V 
Rd ,max   450  0.5bw z fcd --------------------------------------- (4.11)

V 
Rd ,max   21.80  0.34483bw z fcd ---------------------------------- (4.12)
TIP: -Use Table 4.1 to calculate these values easily.
4.2) If VEd > (VRd,max)θ=45°, the beam size should be increased.
4.3) If VEd ≤ (VRd,max)θ=21.8°, then use θ = 21.8° (i.e. cotθ=2.5) and the shear rebar area is;
Asw VEd
 ----------------------------------------------- (4.13)
S 2.5 zf ywd
4.4) If (VRd,max)θ=21.8° < VEd < (VRd,max)θ=45°, the angle θ should be determined first using;
 2VEd 
  0.5  sin 1   ------------------------------------- (4.14)
 bw z fcd 
and then calculate the shear rebar by substituting this angle θ in to;
Asw VEd
 ----------------------------------------- (4.15)
S zf ywd cot 
5) Calculate and check minimum shear reinforcement using equation 4.8.
6) Calculate Spacing and check maximum spacing.

Table 4.1: Values of VRd,max/bwd.

Unity University 14 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

BOND, ANCHORAGE, AND SPLICING OF REINFORCEMENT


4.5 Bond
‘Bond’ in reinforced concrete refers to the adhesion between the reinforcing steel and the
surrounding concrete. It is this bond which is responsible for the transfer of axial force from a
reinforcing bar to the surrounding concrete, thereby providing strain compatibility and
‘composite action’ of concrete and steel. If this bond is inadequate, ‘slipping’ of the reinforcing
bar will occur, destroying full ‘composite action’. Hence, the fundamental assumption of the
theory of flexure, viz. plane sections remain plane even after bending, becomes valid in
reinforced concrete only if the mechanism of bond is fully effective.
In order for reinforced concrete to behave as intended, it is essential that bond stresses be
developed on the interface between concrete and steel, such as to prevent significant slip from
occurring at that interface. If this bond could not be developed, then the bars pull out of the
concrete and the tension drops to zero.
Bond resistance in reinforced concrete is achieved through the following mechanisms:
1) Chemical adhesion: - due to a gum-like property in the products of hydration (formed
during the making of concrete).
2) Frictional resistance: - due to the surface roughness of the reinforcement and the grip
exerted by the concrete shrinkage.
3) Mechanical interlock: - due to the surface protrusions or ‘ribs’ (oriented transversely to the
bar axis) provided in deformed bars.
The contribution from mechanical interlock is significant when compared to the other two.
Evidently, the resistance due to mechanical interlock is not available when plain bars are used.
For this reason, many codes prohibit the use of plain bars in reinforced concrete; except for
lateral spirals, and for stirrups and ties smaller than 10 mm in diameter. For this reason,
deformed bars are now universally used. With such bars, the shoulders of the projecting ribs
bear on the surrounding concrete and result in greatly increased bond strength.
4.6 Bond Stress
There are two types of bond stresses;
 Flexural bond and
 Anchorage bond or development bond.

Figure 4.4: Forces and stresses acting on elemental length of beam.

Unity University 15 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

Flexural Bond
Flexural bond is that which arises in flexural members on account of shear or a variation in
bending moment, which in turn causes a variation in axial tension along the length of a
reinforcing bar. Figure 4.4 shows forces in an isolated piece of a beam of length dx. As shown
in the figure, the moment at one end will generally differ from that at the other end by a small
amount dM. since concrete does not resist tensile stresses, the change in bar force, dT, becomes;
dM
dT  ----------------------------------------------- (4.16)
z
As shown in Fig. 4.4b, this force is resisted by the bond at the contact surface between the rebar
and concrete. Summing horizontal forces,

 udx  P  dT ----------------------------------------------- (4.17)


Where: u = local average bond stress per unit of bar surface area.
ΣP = sum of perimeters of all the bars.
dT dM V
u   --------------------------------- (4.18)
(dx)P z(dx)P zP
Anchorage bond (Development bond)
Anchorage bond or development bond is that which arises over the length of anchorage
provided for a bar or near the end (or cut-off point) of a reinforcing bar; this bond resists the
‘pulling out’ of the bar if it is in tension, or conversely, the ‘pushing in’ of the bar if it is in
compression. This situation is depicted in the cantilever beam of Figure 4.5b, where it is seen
that the tensile stress in the bar segment varies from a maximum fs at the continuous end D to
practically zero at the discontinuous end C.

Figure 4.5: Anchorage bond stress.


The bending moment, and hence the tensile stress fs, are maximum at the section at D.
Evidently, if a stress fs is to be developed in the bar at D, the bar should not be terminated at D,
but has to be extended (‘anchored’) into the support (column) by a certain length CD. At the
discontinuous end C of the bar, the stress is zero. The difference in force between C and D is

Unity University 16 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

transferred to the surrounding concrete through anchorage bond. From equilibrium of forces,
the average anchorage bond stress (as shown in Figure 4.5c) will be;
 fs
As f s   L  uav  uav  --------------------------------- (4.19)
4L
In this equation, L is the length that the bar should be extended to anchor or develop the stress
fs safely. In other words, L is the length of embedment necessary to anchor or develop the full
tensile strength of the bar in order to avoid bond failure. For this reason, this length is called
anchorage length or development length.
Bond Failure Mechanisms
Ultimate bond failures for bars in tension are of two types: the first is direct pullout of the bar,
which occurs when ample confinement is provided by the surrounding concrete. The second
type of failure is splitting of the concrete along the bar when cover, confinement or bar spacing
is insufficient to resist the lateral concrete tension resulting from the wedging effect of the bar
deformations. The latter if more common than the former.

Figure 4.6: Splitting of concrete along reinforcement


4.7 Splicing of Reinforcement
Because of restraints such as transportation and fixing, the length of reinforcement bars is
restricted. For this reason, bars are usually joined or spliced. Splices are required when bars
placed short of their required length (due to non-availability of longer bars) need to be
extended. Splices are also required when the bar diameter has to be changed along the length
(as is sometimes done in columns). The purpose of ‘splicing’ is to transfer effectively the axial
force from the terminating bar to the connecting (continuing) bar with the same line of action
at the junction. This invariably introduces stress concentrations in the surrounding concrete.
These effects should be minimized by:
 using proper splicing techniques;
 keeping the splice locations away from sections with high flexural/shear stresses; and
 Staggering the locations of splicing in the individual bars of a group.
Splicing is generally done in one of the following three ways:
1. Lapping of bars (lap splice)
2. Welding of bars (welded splice)
3. Mechanical connections.

Unity University 17 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

4.8 Provisions of ES EN 2 on Anchorage and Splicing of Reinforcement


4.8.1 Anchorage of Reinforcement
All reinforcement should be anchored so that the forces in them are safely transmitted to the
surrounding concrete by bond without causing cracks or spalling. Longitudinal bars are
anchored by providing sufficient anchorage length in the case of a straight bar. The length
required can be reduced by providing a standard bend or a standard hook or by a standard loop.
Anchorage length can also be reduced by welding a transverse bar. The common methods of
anchorage of longitudinal bars are shown in Figure 4.7.

Figure 4.7: Methods of anchorage for longitudinal rebar.


The anchorage of links and shear reinforcement should normally be achieved by means of
bends and hooks, or by welded transverse reinforcement. A bar should be provided inside a
hook or bend. The of anchorage of links and shear reinforcement are shown in Figure 4.8. Note
that when anchorage is provided by welded transverse reinforcement as shown in Fig. 4.8c and
4.8d, the cover should not be less than either 3ϕ or 50 mm.

Figure 4.8: Anchorage for links.

Unity University 18 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

Ultimate Bond Stress


The ultimate bond resistance shall be sufficient to prevent bond failure. The design value of
the ultimate bond stress, fbd, for ribbed bars may be taken as:
fbd  2.25
1 2 f ctd --------------------------------------- (4.20)

Where
fctd = design tensile strength of concrete.
η1 = a coefficient related to the quality of the bond condition and has a value of;
η1 = 1 for ‘good’ bond conditions (see Figure 4.9 for definition) and
η1 = 0.7 for all other cases.
η2 = a coefficient related to the bar diameter and has a value of;
η2 = 1 if ϕ ≤ 32mm
η2 = (132- ϕ)/100 if ϕ > 32mm

Figure 4.9: Description of bond conditions


Table 4.2 gives the values of the bond stress fbd, for different concrete grades and rebar sizes
assuming ‘good’ bond condition.

Table 4.2: Values of fbd.


Basic anchorage length
The basic required anchorage length, lb,rqd, for anchoring the force Asfyd in a bar is given by;
  sd
lb,rqd  ------------------------------------------- (4.21)
4 fbd
Where σsd is the design stress of the bar. If σsd = fyd, then;
 f yd
lb,rqd  ------------------------------------------- (4.22)
4 fbd

Unity University 19 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

Design anchorage length


The design anchorage length, lbd, is given by;
l b ,rqd  α1 α 2 α 3 α 4 α 5
lbd   ------------------------------------ (4.23)
l b ,min
Where,
lb,min = the minimum anchorage length given by;
For bars in tension: lb ,min  max  0.3l b, rqd ;10 ;100mm

For bars in compression: lb ,min  max  0.6l b,rqd ;10 ;100mm


α1, α2, α3, α4 & α5 are coefficients given in Table 4.3 and takes in to account the influence of;
α1 = the shape of bar α2 = the concrete cover
α3 = confinement by transverse bars α4 = confinement by welded transverse bars
α5 = confinement by transverse pressure.
Note that the product α2∙α3∙α5 should be ≥ 0.7.

Table 4.3: Values of α coefficients.

Figure 4.10: Values of K for beams and slabs.

Unity University 20 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

Figure 4.11: Definition of Cd.


As a simplified alternative to equation 4.23, the tension anchorage of certain shapes shown in
Figure 4.7 may be provided as an equivalent anchorage length, lb,eq. The equivalent anchorage
length shown in Figure 4.7 can be taken as:
For bends, hooks and loops lb,eq = α1∙lb,rqd
For welded transverse bar lb,eq = α4∙lb,rqd
Instead of calculating the design anchorage length as discussed above, the following table can
be used to get the values of the design anchorage length in bar diameters for different
combinations of concrete and reinforcement grades.
This table is prepared based on the following assumption.
 α1= α2 = α4 = α5 = 1 and α3 = 0.9
 ‘Good’ bond condition exists.
 The rebar size is not greater than 32 (ϕ ≤ 32mm).

Table 4.4: Values of Lbd/ϕ.


Therefore, based on Table 4.4, the value of the design anchorage length will be;
A 
lbd  Value from Table     s ,cal  ---------------------------------- (4.24)
A
 s , pro 
Where
As,cal = Theoretical area of reinforcement required by the design
As,pro = Area of reinforcement actually provided
If applicable, modify the result from Equation 4.24 depending on the following conditions.
 If the bond condition is ‘Poor’; divide the value by 0.7
 If ϕ > 32 divide the value by a factor (132- ϕ)/100
 For slabs; divide the value by 0.9 (because α3 = 1 for slabs)

Unity University 21 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

4.8.2 Laps
According to ES EN 2,
 Laps should not be located in areas of high moments /forces.
 Laps between bars should be staggered.
 Laps at any section should be arranged symmetrically.
 Bars in compression and secondary reinforcement may be lapped in the same location.
The arrangement of lapped bars should comply with Figure 4.12. That is;
1) The clear transverse distance between two lapped bars should not be greater than 4ϕ or
50 mm. If it is greater than these, then the lap length should be increased by the amount
by which the limitation is exceeded.
2) The longitudinal distance between two adjacent laps should not be less than 0.3 times
the lap length, l0. Otherwise, the bars should be considered as being lapped in one section.
3) In case of adjacent laps, the clear distance between adjacent bars should not be less than
2ϕ or 20 mm.
When the lapped bars comply with the provisions above, the permissible percentage of lapped
bars in tension may be 100% where the bars are all in one layer. Where the bars are in several
layers the percentage should be reduced to 50%.

Figure 4.12: Arranging adjacent lapping bars.


Lap length
The design lap length, l0, is given by;
lb ,rqd  α1α2α3α5α6
l0   --------------------------------------- (4.25)
l0,min
Where,
l0,min = the minimum lap length and is given by;
l0,min  max  0.3 6l b ,rqd ;15 ; 200mm
α1, α2, α3, & α5 = defined before in Table 4.3; however, for the calculation of α3, ΣAst,min should
be taken as As(σsd / fyd), where As = area of one lapped bar.
α6 = factor dealing with the percentage of lapped reinforcement at a section and is given by;
1
6  but not exceeding 1.5 nor less than 1.
25
ρ1 = percentage of reinforcement lapped within 0.65∙l0 from the center of the lap length
considered (see Figure 4.13).
Values of α6 are given in Table 4.5 below.

Unity University 22 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

Table 4.5: Values of α6.

Figure 4.13: Example for calculation of percentage of lapped bars in one lapped section.
Instead of calculating the design lap length as discussed above, the following tables can be used
to get the values of the design lap length in bar diameters.
Table 4.6 is for beams and it is prepared based on the following assumption.
 α1= α2 = α4 = α5 = 1 and α3 = 0.9
 not more than 33% of the bars are lapped at one place, α6=1.15
 ‘Good’ bond condition exists.
 The rebar size is not greater than 32 (ϕ ≤ 32mm).

Table 4.6: Values of L0/ϕ for beams.


Table 4.7 is for slabs and it is prepared based on the following assumption.
 α1= α2 = α3 = α4 = α5 = 1
 not more than 33% of the bars are lapped at one place, α6=1.15
 ‘Good’ bond condition exists.
 The rebar size is not greater than 32 (ϕ ≤ 32mm).

Table 4.7: Values of L0/ϕ for slabs.

Unity University 23 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

Table 4.8 is for columns and it is prepared based on the following assumption.
 α1= α2 = α4 = α5 = 1 and α3 = 0.9
 More than 50% of the bars are lapped at one place, α6=1.5
 ‘Good’ bond condition exists.
 The rebar size is not greater than 32 (ϕ ≤ 32mm).

Table 4.8: Values of L0/ϕ for columns.


If applicable, modify the values from Table 4.6, 4.7 and 4.8 depending on the following
conditions.
 If the bond condition is ‘Poor’; divide the value by 0.7
 If ϕ > 32 divide the value by a factor (132- ϕ)/100
4.8.3 Transverse reinforcement in the lap zone
Transverse reinforcement is required in the lap zone to resist transverse tension forces.
For bars in tension
 If the diameter of the lapped bars, ϕ < 20 mm, or the percentage of lapped bars in any one
section is less than 25%, then any transverse reinforcement or links necessary for other
reasons may be assumed sufficient for the transverse tensile forces without further
justification.
 If the diameter of the lapped bars, ϕ ≥ 20 mm, the transverse reinforcement should have a
total area, ΣAst ≥ Area of one lapped bar as shown in Fig 4.14(a).
For bars in compression
In addition to the rules for bars in tension one bar of the transverse reinforcement should be
placed outside each end of the lap length and within 4ϕ of the ends of the lap length.

Figure 4.14: Transverse reinforcement in lap zone (a) tension; (b) compression.

Unity University 24 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

4.9 Curtailment of Reinforcements


For economic reasons, it is a common practice to cutoff rebars where they are no longer needed
and the remaining bars are adequate to carry the tensile stress. When a bar is curtailed in a
flexural member, it should be anchored beyond the point when it is no longer required, for a
length of lbd.
The location of points where bars are cutoff or bent is a function of the tension due to moment
and shear. To determine these points, the tensile force diagram or M/Z diagram shall be used.
The forces due to shear may be accommodated by the ‘shift rule’, i.e. by shifting the location
where a bar is no longer required for moment by an amount al in the direction of reducing
moment.
A practical way of using the shift rule is:
A) Determine where the bar can be curtailed based on bending moment alone, and
B) Anchor this bar beyond this location for a distance lbd + a1.
This procedure is illustrated in Figure 4.14. Note that the value of a1 is;
For members with shear reinforcement (e.g. beams) a1 = 0.5ꞏzꞏcotθ = 0.45dꞏcotθ
For members without shear reinforcement (e.g. slabs) a1 = d

Figure 4.15: Illustration of curtailment of longitudinal reinforcement

Unity University 25 Compiled by: Feysel N.


Department of Civil Engineering
RC-I Design of Beams for Shear

4.9.1 Simplified Curtailment Rules


In the absence of explicit calculation, the simplified curtailment rules shown in the figure below
can be used.

Figure 4.16: Simplified curtailment rule for simple beam.

Figure 4.17: Simplified curtailment rule for isolated cantilever beam.

Figure 4.18: Simplified curtailment rule for continuous beam.

Unity University 26 Compiled by: Feysel N.


Department of Civil Engineering

You might also like