You are on page 1of 86

Reinforced Concrete III Dr.

Nasr Abboushi

CHAPTER 13 BEARING, RETAINING AND SHEAR WALLS.

13.1 INTRODUCTION

Definitions—Walls and Wall Loadings


ACI Code Section 2.2 defines a wall as follows:
“Wall—Member, usually vertical, used to enclose or separate spaces.”
This definition fails to consider the structural actions of walls. ACI Section 2.1 also defines
the term “structural walls”:
“Structural wall — Wall proportioned to resist combinations of shears, moments, and axial
forces. A shear wall is a structural wall.”
Major factors that affect the design of structural walls include the following:
a) The structural function of the wall relative to the rest of the structure.
 The way the wall is supported and braced by the rest of the structure.
 The way the wall supports and braces the rest of the structure.
a) The types of loads the wall resists.
b) The location and amount of reinforcement.
Two frequent characteristics of walls are their slenderness, height to thickness ratio, which is
generally higher than for columns, and the reinforcement ratios, generally about a fifth to a
tenth of those in columns.

Types of Walls
Structural walls can be classified as:
a) Bearing walls — walls that are laterally supported and braced by the rest of the
structure that resist primarily in-plane vertical loads acting downward on the top of
the wall (see Figure a below). The vertical load may act eccentrically with respect to
the wall thickness, causing weak-axis bending.
b) Shear walls — walls that primarily resist lateral loads due to wind or earthquakes
acting on the building are called shear walls or structural walls. These walls often
provide lateral bracing for the rest of the structure. (See Figure b below). They resist
gravity loads transferred to the wall by the parts of the structure tributary to the
wall, plus lateral-loads (lateral shears) and moments about the strong axis of the wall.
c) Nonbearing walls — walls that do not support gravity in-plane loads other than their
own weight. These walls may resist shears and moments due to pressures or loads
acting on one or both sides of the wall. Examples are basement walls and retaining
walls used to resist lateral soil pressures. (See Figures c and d).

One-Way and Two-Way Walls


Walls may be supported and restrained against lateral deflections along one to four sides.
Cantilever retaining walls are generally supported solely along the lower edge of the wall.
Such walls act as vertical flexural cantilevers that resist lateral loads from the adjoining soil.
462
Reinforced Concrete III Dr. Nasr Abboushi

Bearing walls are generally laterally supported and restrained against deflection along two
opposite sides, usually the top and bottom supports. Cantilever retaining walls and bearing
walls transfer load in one direction: to supports at the top and bottom of the wall, for
example, or to supports at the east and west edges. In the terminology used for one-way
and two-way slabs, these are referred to as one-way walls. One-way walls may be designed
as wide columns spanning between the top and bottom supports, using ACI Code Chapters
10 and 11, or they may be designed using ACI Code Chapter 14. Walls that transfer load in
more than one direction are called two-way walls. Walls supported on three sides may occur
in open-topped, rectangular tanks; in storage bins for bulk materials; or in counterfort
retaining walls (see the above Figure d). Walls supported on four sides are used to resist
forces or pressures applied perpendicular to the walls.

Minimum reinforcement, wall thickness.


Minimum vertical and horizontal reinforcement shall be in accordance with 14.3.2 and
14.3.3 unless a greater amount is required for shear by 11.9.8 and 11.9.9.
463
Reinforced Concrete III Dr. Nasr Abboushi

14.3.2 — Minimum ratio of vertical reinforcement area to gross concrete area, , shall be:
(a) for deformed bars not larger than with not less than ; or
(b) for other deformed bars; or
(c) for welded wire reinforcement not larger than MW200 or MD200.
14.3.3 — Minimum ratio of horizontal reinforcement area to gross concrete area, , shall
be:
(a) for deformed bars not larger than with not less than ; or
(b) for other deformed bars; or
(c) for welded wire reinforcement not larger than MW200 or MD200.

Walls more than thick, except basement walls, shall have reinforcement for each
direction placed in two layers parallel with faces of wall in accordance with the following
(14.3.4):
(a) One layer consisting of not less than one-half and not more than two-thirds of
total reinforcement required for each direction shall be placed not less than
nor more than one-third the thickness of wall from the exterior surface;
(b) The other layer, consisting of the balance of required reinforcement in that
direction, shall be placed not less than nor more than one-third the
thickness of wall from the interior surface.
Spacing of vertical and horizontal reinforcement shall not exceed nor three times
the wall thickness (14.3.5).
According to 14.3.6, lateral ties for vertical reinforcement are not required as long as the
vertical reinforcement is not required as compression reinforcement or the area of vertical
reinforcement does not exceed times the gross concrete area.
A minimum of two bars in walls having two layers of reinforcement in both directions
and one bar in walls having a single layer of reinforcement must be provided around all
window and door and similar sized openings, with minimum bar extension beyond the
corner of opening equal to the bar development length (14.3.7).
Thickness of nonbearing walls shall not be less than , nor less than the least
distance between members that provide lateral support (14.6.1).

13.2 BEARING WALLS.

Walls used primarily to support gravity loads in buildings are referred to as bearing walls.
Design is by ACI Code Section 14.5, which was derived specifically to apply to walls subjected
to axial loads and moments due to the axial loads acting at an eccentricity of one-sixth of the
thickness of the wall from the midplane of the wall (i.e., at the kern of the wall). The
resulting moments are referred to as weak-axis bending moments. ACI Code Section 14.4
allows the design of bearing walls to be carried out either by:
1. Using the one-way column design and slenderness requirements in ACI Code Sections
10.2, 10.3, 10.10, 10.14, 14.2, and 14.3, or
2. The so-called empirical design method in ACI Code Section 14.5.
464
Reinforced Concrete III Dr. Nasr Abboushi

Walls with strong-axis moments and significant in-plane shear forces acting parallel to the
wall, referred to as shear walls or structural walls, are not covered by ACI Chapter 14,
although the code does not state this. Shear walls will be discussed later.

It is not within the scope of this section to include design aids for a broad range of wall and
loading conditions. The intent is to present examples of various design options and aids. The
designer can, with reasonable effort, produce design aids to fit the range of conditions
usually encountered in practice. For example, strength interaction diagrams such as those
plotted in Fig. (a) ( ) and Fig. (b) ( ) can be helpful design aids for
evaluation of wall strength.

465
Reinforced Concrete III Dr. Nasr Abboushi

Section 14.5 contains the Empirical Design Method which applies to walls of solid
rectangular cross-section with resultant loads for all applicable load combinations falling
within the middle third of the wall thickness at all sections along the height of the wall.
Minimum thicknesses of walls designed by this method are contained in 14.5.3. Walls of
nonrectangular cross-section, such as ribbed wall panels, must be designed by the provisions
of 14.4, or if applicable, 14.8.
Section 14.8 contains the provisions of the Alternate Design Method, which are applicable
to simply supported, axially loaded members subjected to out-of-plane uniform lateral loads,
with maximum moments and deflections occurring at mid-height. Also, the wall cross-
section must be constant over the height of the panel. No minimum wall thicknesses are
prescribed for walls designed by this method.
All walls must be designed for the effects of shear forces. Section 14.2.3 requires that the
design for shear must be in accordance with 11.9 special shear provisions for walls. The
required shear reinforcement may exceed the minimum wall reinforcement prescribed in
14.3.

13.2.1 Axial-Load Capacity (Empirical Design Method)

(ACI Eq. 14-1) was derived in a two-step procedure. First, the capacity of a short wall was
derived. Then this was multiplied by a factor reflecting the effects of slenderness on the
axial-load capacity. The largest eccentricity at which a load can be applied to a plain concrete
wall without developing tensile stresses is at one-sixth of
the wall thickness from the midthickness of the wall (at
the kern point of the section). This load case can be
approximated by a rectangular stress block extending
from the compressed face of the wall for a distance of
two-thirds of the thickness of the wall as shown in the
Figure here. The force per horizontal length of wall, is

This was rounded off to and then multiplied


by the term in the square brackets in the following
Equation to account for the slenderness of the wall. The
slenderness term was derived to give reasonable
agreement with the slenderness effects in ACI Code
Section 10.10. The equation for the axial-load capacity of
a bearing wall is

* ( ) +

where
is the horizontal length of a wall,
is the overall thickness of a wall
466
Reinforced Concrete III Dr. Nasr Abboushi

length of compression member (column or wall) in a frame, measured from center to


center of the joints in the frame.
is the effective length factor for a wall, taken as
 For walls braced top and bottom against lateral translation and
(c) Restrained against rotation at one or both ends (top, bottom, or both)................ 0.8
(d) Unrestrained against rotation at both ends……………………………………………………...... 1.0
 For walls not braced against lateral translation………………………….………..…………....... 2.0
is the strength-reduction factor for compression-controlled sections, taken equal to
0.65.

The next Figure shows typical axial load-moment strength curves for
walls with and The curves yield eccentricity factors (ratios of
strength under eccentric loading to that under concentric loading) of and
for the walls with and .

Thickness, Reinforcement, and Sustained Loads — the above Equation is not affected by the
amount of wall reinforcement and does not allow for creep under sustained axial loads. This
is in contrast to design by ACI Code Chapter 10.
ACI Code Section 14.5.3.1 limits the minimum thicknesses of walls designed using the so-
called empirical design method to of the unsupported height or length of the wall,
whichever is shorter, but not less than . Thickness of exterior basement walls and
foundation walls shall not be less than (14.5.3.2). ACI Code Chapter 14 does not
require wall reinforcement to be designed for the loads on the wall. Instead, ACI Code
Sections 14.3.2 and 14.3.3 give the minimum vertical and horizontal reinforcement ratios.

467
Reinforced Concrete III Dr. Nasr Abboushi

These reinforcement ratios can be written in terms of the maximum spacing of the bars.
Thus, for or smaller bars with not less than , the maximum horizontal and
vertical spacings are as follows:
Vertical steel:

Horizontal steel:

If the reinforcement is in two layers, is the total area of vertical bars within the spacing
, and similarly for the horizontal bars.

ACI Code Section 14.3 requires more reinforcement horizontally than vertically. This reflects
the greater chance that vertical cracks in walls might form as a result of restrained horizontal
shrinkage or temperature stresses, compared with a lower chance that horizontal cracks will
form as a result of restrained vertical stresses. Generally, if shrinkage occurs in the vertical
direction, the shrinkage stresses are dissipated by vertical compression stresses in the wall.

Example
A wall with a vertical height between lateral supports of and a horizontal length of
between intersecting walls supports a uniformly distributed factored gravity load of
, including the self-weight of the wall. The wall is supported on a strip footing
that prevents lateral movement of the bottom of the wall. The wall supports a wooden-
frame roof deck, which acts as a diaphragm to restrain lateral displacement of the top of the
wall. Is a 200-mm-thick wall adequate if ? If so, select reinforcement for the
wall.

Solution:
1. Check whether the Wall Thickness is Sufficient. ACI Code Section 14.5.3.1 limits the
thickness of walls designed by the empirical design method to the larger of .
and of the shorter of the unsupported height or the length. Thus, the minimum
thickness is ( ). A 200-mm-thick wall satisfies the minimum
thickness given in ACI Code Section 14.5.3.1. Use a 200-mm-thick wall.

2. Compute the Capacity of a 1-m-Wide Strip of Wall.

* ( ) +

ACI Code Section 14.5.2 gives for the end restraints described in the statement of
the problem. Walls will seldom have spiral reinforcement. As a result, the wall is an “other”
type of member, and ACI Code Section 9.3.2.2 specifies . We thus have

468
Reinforced Concrete III Dr. Nasr Abboushi

* ( ) +

This value exceeds the applied factored gravity load of ; thus, the wall has
adequate capacity.

3. Select Reinforcement. ACI Code Sections 14.3.2 and 14.3.3 require minimum areas of
and for vertical and horizontal reinforcement, respectively. ACI
Code Section 14.3.4 allows the reinforcement to be placed in one layer, or “curtain”
because the wall thickness is less than . ACI Code Section 14.3.5 gives the
maximum bar spacing parallel to the wall as the smaller of 3 times the wall thickness
— — and an upper limit of . The maximum
spacing of the reinforcement is as follows:

Horizontal spacing of vertical reinforcement—from ACI Code Section 14.3.2:

If the required vertical steel is placed in a single layer of vertical bars,


and the spacing is

Vertical spacing of horizontal reinforcement—from ACI Code Section 14.3.3:

Using a single layer of bars, the spacing of the minimum horizontal reinforcement is

Because the vertical reinforcement is not specifically used in the strength design, ACI Code
Section 14.3.6 does not require the vertical bars to be tied in nonseismic regions provided
that
(a) the area of vertical steel is less than 0.01 times the gross area of the wall, or
(b) the steel is not used as compression steel.
The vertical steel provided has ( ) times the gross area of
the wall.
It is good practice to provide a bar vertically at each end of each curtain of wall steel.
Use a 200-mm-thick wall with , with one curtain of bars at the center of the
wall with vertical Grade-420 bars at o.c. and horizontal Grade-420 bars
at o.c. Add one vertical bar at each end of the wall.

469
Reinforced Concrete III Dr. Nasr Abboushi

13.2.2 Alternate design of slender walls.

According to 14.8.1, the provisions of 14.8 are considered to satisfy 10.10 when flexural
tension controls the out-of-plane design of a wall. The following limitations apply to the
alternate design method (14.8.2):
1. The wall panel shall be simply supported, axially loaded, and subjected to an out-of-
plane uniform lateral load. The maximum moments and deflections shall occur at the
mid-height of the wall (14.8.2.1).
2. The cross-section is constant over the height of the panel (14.8.2.2).
3. The wall cross sections shall be tension-controlled (14.8.2.3).
4. Reinforcement shall provide a design moment strength greater than or equal to
, where is the moment causing flexural cracking due to the applied lateral and
vertical loads. Note that shall be obtained using the modulus of rupture given
by Eq. (9-10) (14.8.2.4).
5. Concentrated gravity loads applied to the wall above the design flexural section shall
be distributed over a width equal to the lesser of
a) the bearing width plus a width on each side that increases at a slope of 2 vertical to 1
horizontal down to the design flexural section or
b) the spacing of the concentrated loads. Also, the distribution width shall not extend
beyond the edges of the wall panel (14.8.2.5) (see next Figure).

6. The vertical stress at the mid-height section shall not exceed (14.8.2.6).

When one or more of these conditions are not satisfied, the wall must be designed by the
provisions of 14.4 (walls subject to axial load or combined flexure and axial load shall be designed
as compression members in accordance with provisions of 10.2, 10.3, 10.10, 10.11, 10.14, 14.2, and
14.3.).

470
Reinforced Concrete III Dr. Nasr Abboushi

According to 14.8.3, the design moment strength for combined flexure and axial loads
at the mid-height cross-section must be greater than or equal to the total factored moment
at this section. The factored moment includes effects and is defined as follows:

where maximum factored moment at the mid-height section of the wall due to
lateral and eccentric vertical loads, not including effects.
factored axial load.
deflection at the mid-height of the wall due to the factored loads,

( )
vertical distance between supports.
modulus of elasticity of concrete (8.5)
moment of inertia of cracked section transformed to concrete,

( ) ( )

modular ratio of elasticity, (14.8.3)


modulus of elasticity of nonprestressed reinforcement.
area of effective longitudinal tension reinforcement in the wall segment

( )( )

area of longitudinal tension reinforcement in the wall segment


specified yield stress of nonprestressed reinforcement
distance from extreme compression fiber to centroid of longitudinal tension
reinforcement
distance from extreme compression fiber to neutral axis corresponding to the
effective longitudinal Reinforcement
horizontal length of the wall

Note that the equation for includes the effects of the factored axial loads and lateral load
( ), as well as the effects ( ).
Substituting Equations for into Equation for results in the following equation for :

( )
The next Figure shows the analysis of the wall according to the provisions of 14.8 for the
case of additive lateral and gravity load effects.

The design moment strength of the wall can be determined from the following
equation:
( )
where

471
Reinforced Concrete III Dr. Nasr Abboushi

and is determined in accordance with 9.3.2.

In addition to satisfying the strength requirement of , the deflection requirement


of 14.8.4 must also be satisfied. In particular, the maximum deflection due to service
loads, including effects, shall not exceed .
is the maximum moment at midheight of wall due to service lateral and eccentric vertical
loads, including effects.
When ( ) , then

( )

When ( ) , then
* ( ) +* ( ) +
( )
* ( ) +
Where:

Example (Design of Precast Panel by the Alternate Design Method (14.8))


Determine the required vertical reinforcement for the precast wall panel shown below. The
roof loads are supported through the webs of the double T-section beams which are
spaced on center.

472
Reinforced Concrete III Dr. Nasr Abboushi

Design data:
Weight of double T-section beam
Roof dead load
Roof live load
Wind load
Concrete ( )
Reinforcing steel .
Eccentricity of the roof loads about the panel
center line

Solution:

1. Trial wall section.


Try .
Try a single layer of vertical
reinforcement:
( ) at
centerline of wall ( ).
For a wide design strip:

( )

2. Distribution width of interior concentrated loads at mid-height of wall.

473
Reinforced Concrete III Dr. Nasr Abboushi

3. Roof loading per meter width of wall


*( )( )+

*( )( )+

Eccentricity of the roof loads about the panel center line .

4. Factored load combinations at mid-height of wall (see figure on page 77).

a) Load comb. 1: .

( )

√ √

( ) ( )

( )( ) ( )( )

( ) ( )

( ) ( )

Therefore, section is tension-controlled .

474
Reinforced Concrete III Dr. Nasr Abboushi

b) Load comb. 2: .

( )

( )( ) ( )( )

( ) ( )

( ) ( )

Therefore, section is tension-controlled .

c) Load comb. 3: .

( )

( )( ) ( )( )

( ) ( )

475
Reinforced Concrete III Dr. Nasr Abboushi

( ) ( )

Therefore, section is tension-controlled .

d) Load comb. 4: .

( )( ) ( )( )

( ) ( )

( ) ( )

Therefore, section is tension-controlled .

5. Check if section is tension-controlled.

Assume section is tension-controlled

476
Reinforced Concrete III Dr. Nasr Abboushi

 Lc1:
 Lc2:
 Lc3:
 Lc4:

( ) ( )

( ) ( )
Tension controlled section.

6. Determine .

√ √

7. Check design moment strength .

a. Load comb. 1

( ) ( )

b. Load comb. 2

( ) ( )

c. Load comb. 3

( ) ( )
477
Reinforced Concrete III Dr. Nasr Abboushi

d. Load comb. 4

( ) ( )

8. Check vertical stress at mid-height section.


Load comb. 2 governs:

9. Check midheight deflection .


maximum moment at midheight of wall due to service lateral and eccentric vertical
loads, including effects.

( )

( )

For ( ) , then

( )

Since is a function of and is a function of , no closed form solution for is


possible. will be determined by iteration.

( ) ( )

One more iteration:

( )

478
Reinforced Concrete III Dr. Nasr Abboushi

( )

No further iterations are required.

( )
Therefore,

The wall is adequate with vertical reinforcement.

13.3 RETAINING WALLS.

Retaining walls are structural members used to provide stability for soil or other materials
and to prevent them from assuming their natural slope. In this sense, the retaining wall
maintains unequal levels of earth on its two faces. The retained material on the higher level
exerts a force on the retaining wall that may cause its overturning or failure. Retaining walls
are used in bridges as abutments, in buildings as basement walls, and in embankments. They
are also used to retain liquids, as in water tanks and sewage-treatment tanks.

13.3.1 Types of retaining walls.

Retaining walls may be classified as follows (refer to next Figure):

1. Gravity walls usually consist of plain concrete or masonry and depend entirely on
their own weight to provide stability against the thrust of the retained material.
These walls are proportioned so that tensile stresses do not develop in the concrete
or masonry due to the exerted forces on the wall. The practical height of a gravity
wall does not exceed .
479
Reinforced Concrete III Dr. Nasr Abboushi

2. Semigravity walls are gravity walls that have a wider base to improve the stability of
the wall and to prevent the development of tensile stresses in the base. Light
reinforcement is sometimes used in the base or stem to reduce the large section of
the wall.

3. The cantilever retaining wall is a reinforced concrete wall that is generally used for
heights from . It is the most common type of retaining structure because
of economy and simplicity of construction. Various types of cantilever retaining walls
are shown in next figure.

480
Reinforced Concrete III Dr. Nasr Abboushi

4. Counterfort retaining walls higher than develop a relatively large bending


moment at the base of the stem, which makes the design of such walls
uneconomical. One solution in this case is to introduce transverse walls (or
counterforts) that tie the stem and the base together at intervals. The counterforts
act as tension ties supporting the vertical walls. Economy is achieved because the
stem is designed as a continuous slab spanning horizontally between counterforts,
whereas the heel is designed as a slab supported on three sides.

5. The buttressed retaining wall is similar to the counterfort wall, but in this case the
transverse walls are located on the opposite, visible side of the stem and act in
compression. The design of such walls becomes economical for heights greater than
. They are not popular because of the exposed buttresses.

6. Bridge abutments are retaining walls that are supported at the top by the bridge
deck. The wall may be assumed fixed at the base and simply supported at the top.

7. Basement walls resist earth pressure from one side of the wall and span vertically
from the basement-floor slab to the first-floor slab. The wall may be assumed fixed at
the base and simply supported or partially restrained at the top.

481
Reinforced Concrete III Dr. Nasr Abboushi

13.3.2 Earth Pressures on Retaining Walls.

The subject of earth pressures is typically covered in geotechnical or foundation courses. For
those students who have not studied this subject, we present a brief review of the main
topics required for wall design.
Retaining walls are generally subjected to gravity loads and to earth pressure due to the
retained material on the wall. Gravity loads due to the weights of the materials are well
defined and can be calculated easily and directly. The magnitude and direction of the earth
pressure on a retaining wall depends on the type and condition of soil retained and on other
factors and cannot be determined as accurately as gravity loads. The theories and procedure
for determining the soil pressure on retaining walls, the stability of retaining walls and the
effect of dynamic reaction on walls were explained in soil mechanics courses.
The lateral force on a wall varies as the wall undergoes lateral movement. The relationship
between the force and the movement is shown in the next Figure. The ordinate of point A
represent the force on a wall which has been held rigidly in place while a soil backfill is
behind it. This is called Earth pressure at rest where there is no relative movement between
the back fill and the soil. This is also called condition where is the corresponding earth
pressure coefficient. values are around . is referred to as coefficient of earth
pressure at rest.

If the wall moves in the direction away from the backfill, the force decreases and after a
small movement reaches a minimum value at point B. This is called active earth pressure.
The corresponding earth pressure coefficient is . values are around .
If the wall is forced against the backfill, the force between the wall and the fill increases,
reaching a maximum value at C. This is called passive earth pressure. The corresponding
earth pressure coefficient . values are around .
The earth pressures and hence the earth pressure coefficients are evaluated using several
earth pressure theories (Taylor, 1964; Bowles, 1996; Das, 2002). The most popular among
them are Rankine’s theory and Coulomb’s theory, which are discussed below. They furnish
482
Reinforced Concrete III Dr. Nasr Abboushi

expressions for active and passive pressures and thrusts caused by a soil mass which is not
subject to seepage forces. Each applies to the cross section of a long wall of constant section
and gives results per unit of running length.
Theoretical soil pressures associated with cohesionless soils are predicted by the Rankine
and Coulomb theories. Although these theories are limited to cohesionless soils, they
provide insight into the pressures other types of soils apply to walls. For cohesive soils that
contain clays or silts, empirical values of soil pressure, based on field measurements, are
available to designers.
Granular materials, such as sand, behave differently from cohesive materials, such as clay, or
from any combination of both types of soils. Although the pressure intensity of soil on a
retaining wall is complex, it is common to assume a linear pressure distribution on the wall.
The pressure intensity increases with depth linearly, and its value is a function of the height
of the wall and the weight and type of soil. The pressure intensity, , at a depth below the
earth's surface may be calculated as follows:

where is the unit weight of soil and is a coefficient that depends on the physical
properties of soil. The value of the coefficient varies from for loose granular soil, such
as sand, to about for cohesive soil, such as wet clay. If the retaining wall is assumed
absolutely rigid, a case of earth pressure at rest develops. Under soil pressure, the wall
may deflect or move a small amount from the earth, and active soil pressure develops, as
shown in Figure below. If the wall moves toward the soil, a passive soil pressure develops.
Both the active and passive soil pressures are assumed to vary linearly with the depth of
wall. For dry, granular, noncohesive materials, the assumed linear pressure diagram is fairly
satisfactory; cohesive soils or saturated sands behave in a different, nonlinear manner.
Therefore, it is very common to use granular materials as backfill to provide an
approximately linear pressure diagram and also to provide for the release or drainage of
water from behind the wall (see page xxxxxx).

483
Reinforced Concrete III Dr. Nasr Abboushi

For a linear pressure, the active and passive pressure intensities are determined as follows:

where and are the approximate coefficients of the active and passive pressures,
respectively.

13.3.3 At-Rest Pressure.

We begin our discussion of soil pressures by considering the forces on a rectangular soil
element, which is located a distance below ground surface (see the figure below). We will
assume that the soil is cohesionless, i.e., a sand or gravel. Soils of this type provide superior
backfill for walls because they exert smaller lateral pressures than soils with a high clay or silt
content. In plan, the element is square with sides 1 unit in length. In this discussion we will
also assume that the water table is located well below the element. (When soil is
submerged, the soil pressures are a function of the buoyant weight—the difference between
the unit weights of soil and water—not the total weight.) Since the vertical pressure
acting on the element is produced by the weight of the column of soil above the element, its
value can be expressed as

where the unit weight of the soil


distance from the ground surface to the top of the element
Experimental studies have shown that the horizontal pressure , which acts on the sides of
the element, can be related to the vertical pressure , by

where , called the at-rest coefficient, ranges from for sands deposited in layers
without mechanical compaction. If the soil is compressed vertically by tamping, the value of
will be higher. If the value of is substituted into Equation for , we can express as

484
Reinforced Concrete III Dr. Nasr Abboushi

Since and are constants, the last equation shows that the lateral pressure varies
linearly with depth below the ground surface. A plot of the variation of the lateral soil
pressure with depth below grade is shown in the previous figure.

Do you have to include At-Rest Pressures for Retaining Wall Design?


An important aspect of "At-Rest" lateral earth pressures is that they typically take place at
zero lateral wall displacement. This means that a wall will experience full "At-Rest" lateral
pressures only if it does not yield. Such a case could take place if a stiff gravity wall fully
bears on bedrock; in such a condition a retaining wall will essentially feel the full "At-rest"
driving soil pressures.
Given that "At-rest" pressures are considerably greater than active earth pressures, one
might conclude that all braced excavations should be designed with at-rest pressures. Doing
so, might actually do greater damage than good. While having a greater capacity might be
beneficial, if a series of supports are prestressed to the "full" theoretical "at-rest" load then
"in-practice" the wall might actually move back into the retained soil causing a series of
unpredicted problems. The author is aware of a diaphragm wall designed this way that
moved as much as back into the retained soil causing severe wall and pavement
cracking in the process. Part of the reason for such observations is that engineers tend to be
on the safe side when providing "at-rest" pressure coefficients and other geotechnical
strength parameters. Thus, the actual "at-rest" coefficient might be smaller than originally
predicted.

13.3.4 Active and Passive Soil pressures.

The lateral soil pressure for which a wall must be designed is somewhat smaller than the at-
rest pressure given by Equation ( ). To establish the lateral soil pressure for
which a wall is designed in practice, we will examine the behavior of the walls in the Figures
below as they are forced into, or move away from, the fill behind the wall. Let us imagine
that the wall in Figure “a” has been inserted into the soil without changing the at-rest state
of stress, and has been braced with struts whose length can be adjusted (that is, the wall can
be moved horizontally in either direction), and that the soil has been excavated from the left
side. If we now imagine that the struts are adjusted so that the wall moves a short distance
outward (away from the backfill), a wedge of soil will be brought to a state of
incipient failure. That is, the soil wedge tends to displace down and to the left, as shown in
Figure on page 86. Along the inclined failure surface ac, shear stresses, acting upward and to
the right, develop to resist the displacement of the soil wedge. These shear stresses are a
function of the normal stresses acting on plane ac. As you can see from considering
equilibrium of the soil wedge in the horizontal direction, the horizontal component of the
shear stresses resists the outward displacement of the soil wedge. As a result of this lateral
restraint, the horizontal pressure the wall must exert on the soil wedge to maintain
equilibrium is smaller than the at-rest pressure. This reduced value of lateral soil pressure

485
Reinforced Concrete III Dr. Nasr Abboushi

is called the active pressure and may be predicted by replacing the coefficient in Equation
for by a new coefficient ; that is

We next imagine that a force , the passive thrust, forces the wall into the backfill a
distance ,. This displacement brings the wedge of soil in Figure “b” to a state of
incipient failure; that is, movement of the soil wedge impends upward and to the right along
plane ac, the lower boundary of the failure wedge. This movement is resisted by (1) the
weight of the soil wedge, and (2) shear stresses along plane ac. Since these shear stresses
oppose the displacement and since the soil wedge is much larger than the wedge associated
with the active case (see Figure a), the magnitude of the passive pressure against the wall
is much larger than the active pressure . The passive pressure, which varies linearly with
depth, is equal to

486
Reinforced Concrete III Dr. Nasr Abboushi

The two theories most commonly used in the calculation of earth pressure are those of
Rankine and Coulomb.

1. In Rankine's approach, the retaining wall is assumed to yield a sufficient amount to


develop a state of plastic equilibrium in the soil mass at the wall surface. The rest of
the soil remains in the state of elastic equilibrium. The theory applies mainly to a
homogeneous, incompressible, cohesionless soil and neglects the friction between
soil and wall. The active soil pressure at a depth h on a retaining wall with a
horizontal backfill based on Rankine's theory is determined as follows:

( )

where

( )

angle of internal friction of the soil (see next Table)


Total active pressure,

( )

The resultant, , acts at from the base (see the above Figure). When the earth is
surcharged at an angle to the horizontal, then


( )

The resultant, , acts at and is inclined at an angle to the horizontal (see next
Figure). The values of expressed by the previous equation for different values of and
are shown in Table below.
487
Reinforced Concrete III Dr. Nasr Abboushi

Passive soil pressure develops when the retaining wall moves against and compresses the
soil. The passive soil pressure at a depth on a retaining wall with horizontal backfill is
determined as follows:

( )

where

( )

Total passive pressure,

( )

The resultant, , acts at from the base (see the Figure on page 87). When the earth is
surcharged at an angle to the horizontal, then


( )

488
Reinforced Concrete III Dr. Nasr Abboushi

acts at and is inclined at an angle to the horizontal (next Figure). The values of
expressed by the above equation for different values of and are shown in next Table.
The values of and vary with the type of backfill used. As a guide, common values of
and are given before.

2. In Coulomb's theory, the active soil pressure is assumed to be the result of the
tendency of a wedge of soil to slide against the surface of a retaining wall. Hence,
Coulomb's theory is referred to as the wedge theory. While it takes into
consideration the friction of the soil on the retaining wall, it assumes that the surface
of sliding is a plane, whereas in reality it is slightly curved. The error in this
assumption is negligible in calculating the active soil pressure. Coulomb's equations
to calculate the active and passive soil pressure are as follows:

where
( )

( ) ( )
( )* √ +
( ) ( )

where
angle of internal friction of soil,

489
Reinforced Concrete III Dr. Nasr Abboushi

angle of the soil pressure surface from the vertical,


angle of friction along the wall surface (angle between soil and concrete),
angle of surcharge to the horizontal.
The total active soil pressure is

When the wall surface is vertical, , and if , then in the previous equation
reduces to equation of Rankine (page 88).
Passive soil pressure is

( )

where
( )

( ) ( )
( )* √ +
( ) ( )

The values of and vary with the type of backfill used. As a guide, common values of
and are given in Table on page 88.

3. When the soil is saturated, the pores of the permeable soil are filled with water,
which exerts hydrostatic pressure. In this case the buoyed unit weight of soil must be
used. The buoyed unit weight (or submerged unit weight) is a reduced unit weight of
soil and equals minus the weight of water displaced by the soil. The effect of the
hydrostatic water pressure must be included in the design of retaining walls
subjected to a high water table and submerged soil. The value of the angle of internal
friction may be used, as shown in Table on page xxx.

13.3.5 Effect of surcharge

Different types of loads are often imposed on the surface of the backfill behind a retaining
wall. If the load is uniform, an equivalent height of soil, , may be assumed acting on the
wall to account for the increased pressure. For the wall shown in figure below, the horizontal
pressure due to the surcharge is constant throughout the depth of the retaining wall.

where
equivalent height of soil, .
pressure of the surcharge, .
unit weight of soil,

The total pressure is


490
Reinforced Concrete III Dr. Nasr Abboushi

( )

In the case of a partial uniform load acting at a distance from the wall, only a portion of the
total surcharge pressure affects the wall (see next figure).

It is a common practice to assume that the effective height of pressure due to partial
surcharge is , measured from point B to the base of the retaining wall. The line AB forms
an angle of with the horizontal.
In the case of a wheel load acting at a distance from the wall, the load is to be distributed
over a specific area, which is usually defined by known specifications such as AASHTO and
AREA specifications.

491
Reinforced Concrete III Dr. Nasr Abboushi

13.3.6 Friction on the retaining wall base.

The horizontal component of all forces acting on a retaining wall tends to push the wall in a
horizontal direction. The retaining wall base must be wide enough to resist the sliding of the
wall. The coefficient of friction to be used is that of soil on concrete for coarse granular soils
and the shear strength of cohesive soils. The coefficients of friction that may be adopted
for different types of soil are as follows:
• Coarse-grained soils without silt,
• Coarse-grained soils with silt,
• Silt,
• Sound rock,
The total frictional force, , on the base to resist the sliding effect is

where
the coefficient of friction,
the vertical force acting on the base,
passive resisting force.
The factor of safety against sliding is

where is the horizontal component of the active pressure, . The factor of safety
against sliding should not be less than 1.5.

13.3.7 Stability against overturning

The horizontal component of the active pressure, , tends to overturn the retaining wall
about the point zero on the toe (see next figure). The overturning moment is equal to
. The weight of the concrete and soil tends to develop a balancing moment, or
lightening moment, to resist the overturning moment. The balancing moment for the case of
the wall shown below is equal to

The factor of safety against overturning is

This factor of safety should not be less than .


The resultant of all forces acting on the retaining wall, , intersects the base at point
(next figure). In general, point does not coincide with the center of the base, , thus
causing eccentric loading on the footing. It is desirable to keep point within the middle
third to get the whole footing under soil pressure. (The case of a footing under eccentric
load was discussed in Chapter 11.)
492
Reinforced Concrete III Dr. Nasr Abboushi

13.3.8 Behavior of a Cantilever Retaining Wall.

Since frictional forces of unknown magnitude develop between the soil and both the sides of
the wall and top of the footing, the forces acting on the wall are indeterminate. If we assume
that rotation of the wall produces a Rankine active state of stress in a V-shaped region
behind the wall (see the cross-hatched region in the next Figure), the stress on a vertical
section passing through the heel can be evaluated for a horizontal and sloping backfills by
eqations in section 13.3.4 (see page 93). Therefore, when analyzing walls we typically work
with a free body consisting of the wall and the soil located directly above the heel.
493
Reinforced Concrete III Dr. Nasr Abboushi

Since concrete has a low tensile strength, reinforcement is required on the tension sides of
flexural members. To establish the position of reinforcement in a cantilever retaining wall,
we will examine the deflected shapes of the three elements that make up the wall, that is,
the stem, heel, and toe (if a key is used, it must also be reinforced). The lateral earth
pressure on the back face of a cantilever retaining wall bends the top of the stem outward.
This displacement creates, on horizontal sections through the stem, vertical compression
stresses adjacent to the front face and tension stresses adjacent to the rear face. The lateral
earth pressure also causes the wall to rotate about the toe. This action lifts the heel of the
footing into the backfill. As a result of the rotational movement, the soil pressure on the
base of the footing reduces under the heel and increases under the toe. Although soil is not
a linear elastic material, designers assume that the soil pressure varies linearly along the
length of the base. Since this assumption has produced many safe structures that have
performed well over time, engineers have adopted this simple soil distribution. The
deflected shape of the wall and footing, created by these-actions, is shown to an
exaggerated scale in the Figure. Free-body diagrams of the stem, heel, and toe show that
each of these members may be treated as a cantilever—hence the name cantilever retaining
wall.
To design the stem, we assume it behaves like a cantilever with a fixed end located at the
top of the foundation (see plane 1 in the previous figure). Vertical flexural steel adjacent to

494
Reinforced Concrete III Dr. Nasr Abboushi

the rear face of the stem is required. The critical section for shear is located a distance
above the base.
Since the heel cantilever supports a large weight of soil above its top surface while the soil
pressure acting upward on the base is relatively small, it bends concave downward.
Therefore, reinforcement is required in the top of the heel. Because vertical support for the
heel is provided by the flexural steel in the back face of the stem, we assume the heel
cantilever is fixed on a vertical plane at this point (see plane 3 in the precious figure).
At the toe of the footing, the bearing pressures on the base (the underside of the footing)
are large while the load applied by the footing weight and the shallow depth of soil covering
the toe is small. These large differences in forces cause the toe of the footing to bend
upward. Accordingly, in this region, reinforcement is required in the bottom of the footing.
Often designers specify that the vertical flexural steel in the wall be carried to the bottom of
the footing, bent , and extended into the toe to serve as the toe reinforcement. As a
conservative measure, at the option of the designer, smaller amounts of steel are often
added to the bottom of the heel cantilever.
The distribution of soil pressure on the base of the foundation is influenced by the point at
which the resultant forces acting on the structure intersect the base of the foundation. If
possible, designers try to proportion walls so that the resultant of the forces acting on the
base falls within the middle third of the base. Resultants that fall outside the middle third
produce large compressive stresses under the base in the region near the toe. If the soil's
bearing capacity is exceeded, these stresses may produce a bearing failure, or in the case of
a compressible soil, excessive tilting of the wall.

13.3.9 Proportions of retaining walls

The design of a retaining wall begins with a trial section and approximate dimensions. The
assumed section is then checked for stability and structural adequacy. The following rules
may be used to determine the approximate sizes of the different parts of a cantilever
retaining wall.
1. Height of the wall: The overall height of the wall is equal to the difference in
elevation required plus to
2. Thickness of the stem: The intensity of the pressure increases with the depth of the
stem and reaches its maximum value at the base level. Consequently the maximum
bending moment and shear in the stem occur at its base. The stem base thickness
may be estimated as to of the height, . The thickness at the top of the stem
may be assumed to be to . Because retaining walls are designed for
active earth pressure, causing a small deflection of the wall, it is advisable to
provide the face of the wall with a batter (taper) of per meter of height, ,
to compensate for the forward deflection. For short walls up to high, a constant
thickness may be adopted.

495
Reinforced Concrete III Dr. Nasr Abboushi

3. Length of the base: An initial estimate for the length of the base of to of the wall
height, , may be adopted.
4. Thickness of the base: The base thickness below the stem is estimated as the same
thickness of the stem at its base, that is, to of the wall height. A minimum
thickness of about is recommended. The wall base may be of uniform
thickness or tapered to the ends of the toe and heel, where the bending moment is
zero.
5. The approximate initial proportions of a cantilever retaining wall are shown in the
figure below.

13.3.10 Design requirements.

The ACI Code, Chapter 14, provides methods for bearing wall design (see page 70). The main
requirements are as follows:
1. The minimum thickness of bearing walls is the supported height or length,
whichever is shorter, but not less than .

496
Reinforced Concrete III Dr. Nasr Abboushi

2. The minimum area of the horizontal reinforcement in the wall is , where


is the gross concrete wall area. This value may be reduced to if
or smaller deformed bars with are used. For welded wire fabric
(plain or deformed), the minimum steel area is .
3. The minimum area of the vertical reinforcement is , but it may be reduced
to if or smaller deformed bars with are used. For
welded wire fabric (plain or deformed), the minimum steel area is .
4. The maximum spacing of the vertical or the horizontal reinforcing bars is the smaller
of or three times the wall thickness.
5. If the wall thickness exceeds , the vertical and horizontal reinforcement
should be placed in two layers parallel to the exterior and interior wall surfaces, as
follows:
• For exterior wall surfaces, at least of the reinforcement (but not more than )
should have a minimum concrete cover of but not more than of the wall
thickness.
This is because the exterior surface of the wall is normally exposed to different
weather conditions and temperature changes.

• For interior wall surfaces, the balance of the required reinforcement in each direction
should have a minimum concrete cover of but not more than of the wall
thickness.

• The minimum steel area in the wall footing (heel or toe), according to the ACI Code,
Section 10.5.4, is that required for shrinkage and temperature reinforcement, which
is when and when or .
Because this minimum steel area is relatively small, it is a common practice to
increase it to that minimum required for flexure:

13.3.11 Drainage

The earth pressure discussed in the previous sections does not include any hydrostatic
pressure. If water accumulates behind the retaining wall, the water pressure must be
included in the design. Surface or underground water may seep into the backfill and develop
the case of submerged soil. To avoid hydrostatic pressure, drainage should be provided
behind the wall. If well-drained cohesionless soil is used as a backfill, the wall can be
designed for earth pressure only. The drainage system may consist of one or a combination
of the following:

497
Reinforced Concrete III Dr. Nasr Abboushi

1. Weep holes in the retaining wall of or more in diameter and spaced


about on centers horizontally and vertically (see next Figure a).
2. Perforated pipe in diameter laid along the base of the wall and
surrounded by gravel (see next Figure b).
3. Blanketing or paving the surface of the backfill with asphalt to prevent seepage
of water from the surface.
4. Any other method to drain surface water.

498
Reinforced Concrete III Dr. Nasr Abboushi

Table Conditions

The next Figure “a” shows a wall of height The groundwater table is located at a depth
below the ground surface, and there is no compensating water on the other side of the wall.
For , the active lateral earth pressure can be given as . (For lateral earth
pressure at rest ).
The variation of , with depth is shown by triangle in the Figure “a”. However, for
(i.e., below the groundwater table), the pressure on the wall is found from the
effective stress and pore water pressure components via the equation

( )

where ( ) the effective unit weight of soil. So the effective active lateral
pressure
[ ( )]

The variation of , with depth is shown by in Figure “a”. Again the lateral pressure
from pore water is
( )

The variation of with depth is shown in Figure “b”.


Hence, the total lateral pressure from earth and water at any depth , is equal to

[ ( )] ( )

In other words, if a portion of the retained height is below a water table, the active pressure
of the saturated soil will increase below that level. This additional pressure for the saturated
soil is equal to the pressure of water, plus the submerged weight of the soil
( ), plus the surcharge of the soil

499
Reinforced Concrete III Dr. Nasr Abboushi

above the water table.The submerged weight of a soil can be approximated as


This pressure diagram is shown in Figure below.

𝑆𝑢𝑟𝑐 𝑎𝑟𝑔𝑒 𝐶𝑎
𝐶𝑎 𝜔 𝐻
𝐶𝑎 𝜔 𝐻 𝐶𝑎 (𝜔𝑠𝑎𝑡 𝜔𝑤 ) 𝐻
𝜔𝑤 𝐻

where

unit weight of water,


unit weight of dry soil,
saturated unit weight of soil,
the effective unit weight of soil, .

Example
The next Figure a shows a - -high retaining wall. The wall is restrained from yielding
(pressure at rest). Calculate the lateral pressure at rest for the shown wall.

500
Reinforced Concrete III Dr. Nasr Abboushi

Given: Weight of backfill is , angle of internal friction is ,


saturated unit weight of soil

Solution:

At

At

At
( ) ( )( )

( ) ( )

501
Reinforced Concrete III Dr. Nasr Abboushi

Example
The trial section of a semigravity plain concrete retaining wall is shown below. It is required
to check the safety of the wall against overturning, sliding, and bearing pressure under the
footing.

Given: Weight of backfill is ,


angle of internal friction is ,
coefficient of friction between concrete and soil is ,
allowable soil pressure is , and

Solution:

1. Using the Rankine equation,

( ) ( )

The passive pressure on the toe is that for a height of , which is small and can be
neglected.

acts at a distance from the bottom of the base.


2. The overturning moment is
3. Calculate the balancing moment, , taken about the toe end :

Weight [ ] Arm [ ] Moment [ ]

∑ ∑

4. The factor of safety against overturning is

502
Reinforced Concrete III Dr. Nasr Abboushi

5. The force resisting sliding, is . The factor of


safety against sliding is
6. Calculate the soil pressure under the base:
503
Reinforced Concrete III Dr. Nasr Abboushi

a. The distance of the resultant from toe end is


The eccentricity is .

The resultant acts just inside the middle third of the base and has an eccentricity of
from the center of the base. For a length of the footing, the effective
length of footing is .
b. The moment of inertia is . Area .
c. The soil pressures at the two extreme ends of the footing are
. The moment is ,
.

7. Check the design strength of concrete section at point of the toe.


a. Soil pressure at (from geometry) is

b. is calculated at due to a rectangular stress and a triangular stress.

( ) ( )

c. The design moment strength of plain concrete is


√ √
where

The section is adequate. No other sections to be checked.

Example
Design a cantilever retaining wall shown below.
Given: Weight of backfill is ,
weight of RC is ,
angle of internal friction is ,
coefficient of friction between concrete and soil is ,
surcharge

504
Reinforced Concrete III Dr. Nasr Abboushi

allowable soil pressure is , and

25 cm Surcharge 𝑝 𝐾𝑁 𝑚
3.5 m

40 cm
50 cm

50 cm
50 cm 40 cm 130 cm

2.2 m

Solution:

𝜔 𝜔

𝑃𝑎
𝜔

Neglect 30 cm
of earth 𝜔

𝑃𝑝 O 𝑃𝑎
505
Reinforced Concrete III Dr. Nasr Abboushi

( ) ( )

In the calculation of the passive pressure, the top of the earth at the toe side is
usually neglected.

The force resisting sliding, is . The


factor of safety against sliding is

The overturning moment is


The balancing moment, , taken about the toe end is
The factor of safety against overturning is

Weight [ ] Arm [ ] Moment [ ]

∑ ∑

Calculate the soil pressure under the base:


The distance of the resultant from toe end is


The eccentricity is .
506
Reinforced Concrete III Dr. Nasr Abboushi

The resultant acts just inside the middle third of the base and has an eccentricity of
from the center of the base.

The moment of inertia is . Area .

The soil pressures at the two extreme ends of the footing are . The moment

is , .

Design of Stem:
 Shear design:

The critical section for shear is at distance from the face of base (bottom of the stem
(wall)). For simplicity, the factored shear force at the bottom of the stem is

√ √

The thickness of at the stem end is adequate enough.

 Flexural design:

[ ]
Take

( )

507
Reinforced Concrete III Dr. Nasr Abboushi

( √ ) ( √ )

Try with

Temperature and shrinkage reinforcement: The minimum horizontal reinforcement at the


base of the wall according to ACI Code, Section 14.3, is

assuming a bars or smaller. .


Because the front face of the wall is mostly exposed to temperature changes, use one-half
to two-thirds of the horizontal bars at the external face of the wall and place the balance at
the internal face.

Use horizontal bars spaced at ( ) at both


the internal and external surfaces of the wall. Use vertical bars spaced at at
the front face of the wall to support the horizontal temperature and shrinkage
reinforcement.

Dowels for the wall vertical bars: The anchorage length of bars into the footing must
be at least . Use an embedment length of into the footing, if the depth of
the footing slab is not enough, try the hook development.
If these dowels are spliced to the vertical stem reinforcing with no more than one-half the
bars being spliced within the required lap length, the splices will fall into the class B category
(ACI Code 12.15) and their lap length should at least equal .
Actually, a much more refined design can be made that involves more cutting of bars. For
such a design, a diagram comparing the theoretical steel area required at various elevations
in the stem and the actual steel furnished is very useful. It is to be remembered (Code
12.10.3) that the bars cut off must run at least a distance or 12 diameters beyond their
theoretical cutoff points and must also meet the necessary development length
requirements.

Design of toe:
The toe of the base acts as a cantilever beam subjected to upward pressures, as calculated
before. The factored soil pressure is obtained by multiplying the service load soil pressure by
a load factor of 1.6, because it is primarily caused by the lateral forces.

508
Reinforced Concrete III Dr. Nasr Abboushi

𝐾𝑁 𝑚

415 mm 𝐴
85 mm

𝐾𝑁 𝑚

𝐾𝑁 𝑚 𝐾𝑁 𝑚
𝐾𝑁 𝑚

The downward pressure due to self-weight of the toe slab

The pressure at the face (section A)

 Shear Design:
The pressure at distance from the face

( )

√ √

The thickness of of the toe slab is adequate enough.

 Flexural design:

( )

Take

509
Reinforced Concrete III Dr. Nasr Abboushi

( )

( √ ) ( √ )

Minimum shrinkage

Temperature and shrinkage reinforcement in the longitudinal direction is not needed in the
heel or toe, but it may be preferable to use minimal amounts of reinforcement in that
direction, say, bars spaced at .

Design of heel:
Load factor of 1.2 is used to calculate the factored bending moment and shear force due to
the backfill and concrete, whereas a load factor of 1.6 is used for the surcharge. The upward
soil pressure is neglected, because it will reduce the effect of the backfill and concrete on
the heel.

𝐾𝑁 𝑚

 Shear Design:
According to normal ACI Code procedures, the first critical section for shear would be a
distance from the face of support. However, the justification for this provision of the ACI
Code is the presence, in the usual case, of vertical compressive stress near a support which

510
Reinforced Concrete III Dr. Nasr Abboushi

tends to decrease the likelihood of shear failure in that region. However, the cantilevered
heel slab is essentially hung from the bottom of the stem by the flexural tensile steel in the
stem, and the concrete compression normally found near a support is absent here.
Consequently, the critical section for shear in the heel slab will be taken at the back face of
the stem.
The downward pressure due to self-weight of the heel slab and soil backfill is
( )

√ √

The thickness of of the heel slab is adequate enough.

 Flexural design:

Take

( )

( √ ) ( √ )

Minimum shrinkage
In the longitudinal direction use

511
Reinforced Concrete III Dr. Nasr Abboushi

Example
Design a cantilever retaining wall to support a bank of earth high. The top of the earth
is to be level with a surcharge of .
Given: The weight of the backfill is , the angle of internal friction is , the
coefficient of friction between concrete and soil is , the coefficient of friction
between soil layers is , allowable soil bearing capacity is , ,
and .

Solution:
1. Determine the dimensions of the retaining wall using the approximate relationships
shown in Figure on page 94.
a. Height of wall: Allowing underground level, the height of the wall becomes
.
b. Base thickness: Assume base thickness is , or take
the base thickness . The height of the stem is .
c. Base length: The base length varies between and . Assuming an
average value of , then the base length equals , say,
. The projection of the base in front of the stem varies between and
. Assume a projection of , say, .

512
Reinforced Concrete III Dr. Nasr Abboushi

d. Stem thickness: The maximum stem thickness is at the bottom of the wall and
varies between and . Choose a maximum stem thickness equal to that
of the base, or . Select a practical minimum thickness of the stem at the
top of the wall of . The minimum batter of the face of the wall is
. For a wall, the minimum batter is , which is
less than the provided. The trial dimensions of the wall
are shown in the previous figure.

2. Using the Rankine equation:

( ) ( )

3. The factor of safety against overturning can be determined as follows:


a. Calculate the actual unfactored forces acting on the retaining wall. First, find
those acting to overturn the wall:
( )

b. The overturning moment is

c. Calculate the balancing moment against overturning (see next figure):

Force [ ] Arm [ ] Moment [ ]

∑ ∑

Factor of safety against overturning

513
Reinforced Concrete III Dr. Nasr Abboushi

4. Calculate the base soil pressure. Take moments about the toe end 0 (next figure) to
determine the location of the resultant of the vertical forces.

∑ ∑

The eccentricity is .

The resultant acts just inside the middle third of the base and has an eccentricity of
from the center of the base.
For a length of the footing, area .
The moment of inertia is .

514
Reinforced Concrete III Dr. Nasr Abboushi

The soil pressures at the two extreme ends of the footing are . The moment
is , .

Soil pressure is adequate.

5. Calculate the factor of safety against sliding. A minimum factor of safety of 1.5 must be
maintained.

In the calculation of the passive pressure, the top of the earth at the toe side is
usually neglected
( )

Force causing sliding


Resisting force is

Factor of safety against sliding is

The resistance provided does not give an adequate safety against sliding. In this case, a key
should be provided to develop a passive pressure large enough to resist the excess force
that causes sliding. Another function of the key is to provide sufficient development length
for the dowels of the stem. The key is therefore placed such that its face is about
from the back face of the stem (see next figure). In the calculation of the passive pressure,
the top of the earth at the toe side is usually neglected, leaving a height of in
this example. Assume a key depth of and a width of .
( ) ( )

The sliding may occur now on the surfaces AC, CD, and EF (see next figure). The sliding
surface AC lies within the soil layers with a coefficient of internal friction
, whereas the surfaces CD and EF are those between concrete and soil with a
coefficient of internal friction of , as given in this example. The frictional resistance is
.

The pressure at point C

( )

515
Reinforced Concrete III Dr. Nasr Abboushi

( ) ( )

The total resisting force is

The factor of safety against sliding is

6. Design the wall (stem).


a. Main reinforcement: The lateral forces applied to the wall are calculated using a load
factor of 1.6. The critical section for bending moment is at the bottom of the wall,
height .
Calculate the applied ultimate forces:
( ) ( )
( ) ( )

516
Reinforced Concrete III Dr. Nasr Abboushi

( )

The total depth used is , , and ,


assuming for bar diameter.
Take

( )

( √ ) ( √ )

Try with
The minimum vertical according to the ACI Code, Section 14.3, is

Because the moment decreases along the height of the wall, may be reduced according
to the moment requirements. It is practical to use one or spacing, for the lower half and a
second , or spacing, for the upper half of the wall. To calculate the moment at midheight
of the wall, from the top.
( ) ( )
( ) ( )

( )

The total depth at midheight of wall is and


, assuming for bar diameter.

( )

517
Reinforced Concrete III Dr. Nasr Abboushi

( √ ) ( √ )

The minimum vertical according to the ACI Code, Section 14.3, is

Try with

Use vertical bars spaced at with similar spacing to the lower vertical steel
bars in the wall.

b. Temperature and shrinkage reinforcement: The minimum horizontal reinforcement


at the base of the wall according to ACI Code, Section 14.3, is

(for the bottom half), assuming a bars or smaller.

(for the upper half) with ( ) .


Because the front face of the wall is mostly exposed to temperature changes, use one-half
to two-thirds of the horizontal bars at the external face of the wall and place the balance at
the internal face.

Use horizontal bars spaced at ( ) at both


the internal and external surfaces of the wall for the bottom half of the wall and with the
same manner for the upper half of the wall. Use vertical bars spaced at at the
front face of the wall to support the horizontal temperature and shrinkage reinforcement.

c. Dowels for the wall vertical bars: The anchorage length of bars into the footing
must be at least . Use an embedment length of into the footing and
the key below the stem.
If these dowels are spliced to the vertical stem reinforcing with no more than one-half the
bars being spliced within the required lap length, the splices will fall into the class B category
(ACI Code 12.15) and their lap length should at least equal .
Actually, a much more refined design can be made that involves more cutting of bars. For
such a design, a diagram comparing the theoretical steel area required at various elevations
in the stem and the actual steel furnished is very useful. It is to be remembered (Code
12.10.3) that the bars cut off must run at least a distance or 12 diameters beyond their
theoretical cutoff points and must also meet the necessary development length
requirements.

518
Reinforced Concrete III Dr. Nasr Abboushi

d. Design for shear: The critical section for shear is at a distance from the
bottom of the stem. At this section, the distance from the top equals
.
( ) ( )
( ) ( )

Total

√ √

The thickness of at the stem end is adequate enough.

7. Design of the heel: A load factor of 1.2 is used to calculate the factored bending
moment and shearing force due to the backfill and concrete, whereas a load factor of
1.6 is used for the surcharge. The upward soil pressure is neglected, because it will
reduce the effect of the backfill and concrete on the heel.
The downward pressure due to self-weight of the heel slab and soil backfill is
( )
The critical section for shear is usually at a distance from the face of the wall when the
reaction introduces compression into the end region of the member. In this case, the critical
section will be considered at the face of the wall, because tension and not compression
develops in the concrete

√ √

The thickness of of the heel slab is not adequate. The section must be increased by
the ratio
Required
Total thickness required
Take , and

( )

519
Reinforced Concrete III Dr. Nasr Abboushi

Take

( )

( √ ) ( √ )

Minimum shrinkage

The development length for the top bars equals . Therefore, the bars must be
extended into the toe of the base.
Temperature and shrinkage reinforcement in the longitudinal direction is not needed in the
heel or toe, but it may be preferable to use minimal amounts of reinforcement in that
direction, say, bars spaced at .

8. Design of the toe: The toe slab acts as a cantilever projecting outward from the face of
the stem. It must resist the upward pressures and the downward load of the toe slab
itself, each multiplied by appropriate load factors. The downward load of the earth fill
over the toe will be neglected because it is subject to possible erosion or removal.
A load factor of 1.6 will be applied to the service load bearing pressures. For the toe
slab, the more severe loading case results from backfill and surcharge on the heel slab.
Because the self-weight of the toe slab tends to reduce design moments and shears, it
will be multiplied by a load factor of 0.9 ( ). The critical section for the
bending moment is at the front face of the stem. The critical section for shear is at a
distance from the front face of the stem, because the reaction in the direction of
shear introduces compression into the toe.
The toe is subjected to an upward pressure from the soil and downward pressure due to
self-weight of the toe slab.
The downward pressure due to self-weight of the toe slab

520
Reinforced Concrete III Dr. Nasr Abboushi

The pressure at the face

The pressure at distance from the face


( )

( ) ( )

√ √

The thickness of of the toe slab is adequate enough.

( )

Take

( )

( √ ) ( √ )

Minimum shrinkage

Use bars spaced at , similar to the heel reinforcement. Development length of


bars equals . Extend the bars into the heel . The final reinforcement
details are shown in the figure below.

521
Reinforced Concrete III Dr. Nasr Abboushi

522
Reinforced Concrete III Dr. Nasr Abboushi

9. Shear keyway between wall and footing: In the


construction of retaining walls, the footing is cast first
and then the wall is cast on top of the footing at a
later date. A construction joint is used at the base of
the wall. The joint surface takes the form of a keyway,
as shown in next figure, or is left in a very rough
condition. The joint must be capable of transmitting
the stem shear into the footing.

10. Proper drainage of the backfill is essential in this design, because the earth pressure
used is for drained backfill. Weep holes should be provided in the wall, in
diameter and spaced at in the horizontal and vertical directions.

13.4 BASEMENT WALLS

It is a common practice to assume that basement walls span vertically between the
basement-floor slab and the first-floor slab. Two possible cases of design should be
investigated for a basement wall.
First, when the wall only has been built on top of the basement floor slab, the wall will be
subjected to lateral earth pressure with no vertical loads except its own weight. The wall in
this case acts as a cantilever, and adequate reinforcement should be provided for a
cantilever wall design. This case can be avoided by installing the basement and the first-floor
slabs before backfilling against the wall.
Second, when the first-floor and the other floor slabs have been constructed and the
building is fully loaded, the wall in this case will be designed as a propped cantilever wall
subjected to earth pressure and to vertical load (fixed-pin supports or pin-pin supports).
The shear and moment diagrams for these support conditions with uniform and triangular
loads are presented in the next figure (can be obtained using any structural software).
In calculations, it is assumed that the backfill is dry, but it is necessary to investigate the
presence of water pressure behind the wall. The maximum water pressure occurs when the
whole height of the basement wall is subjected to water pressure.
The maximum pressure may not be present continuously behind the wall. Therefore, if the
ground is intermittently wet, a percentage of the preceding pressure may be adopted. Water
may be prevented from collecting against the wall by providing drains at the lower end of
the wall.
In addition to drainage, a waterproofing or damp-proofing membrane must be laid or
applied to the external face of the wall.

523
Reinforced Concrete III Dr. Nasr Abboushi

524
Reinforced Concrete III Dr. Nasr Abboushi

Example
Determine the thickness and necessary reinforcement for the basement retaining wall
shown below.
Given: Weight of dry backfill , angle of internal friction , , and
.

Solution:
The wall spans vertically and will be considered as fixed at the bottom end and propped at
the top. Consider a span of , as shown in the above figure. For these data, the
different lateral pressures on a length of the wall are as follows:

( ) ( )

525
Reinforced Concrete III Dr. Nasr Abboushi

( )

 Due to active soil pressure, , and


.
 Due to surcharge, , and
.
is due to triangular loading, whereas is due to uniform loading.
Using moment coefficients of a propped beam subjected to triangular and uniform loads, or
using structural analysis software we obtain the shear and moment diagrams.

Using the moment coefficients

( ) ( )
( ) ( )
Maximum positive bending moment within the span occurs at the section of zero shear.

( )
526
Reinforced Concrete III Dr. Nasr Abboushi

Or in the form
For the positive moment,

* +

Assume wall thickness , assuming for


bar diameter.
Take
For

( )

( √ ) ( √ )

The minimum vertical according to the ACI Code, Section 14.3, is


( )

For

, assuming for bar diameter.

( )

( √ ) ( √ )

527
Reinforced Concrete III Dr. Nasr Abboushi

The minimum vertical according to the ACI Code, Section 14.3, is


( )

Longitudinal reinforcement: Use a minimum steel ratio of (ACI Code, Section 14.3),
or . Use bars spaced at on each
side of the wall.
If the bending moment at the base of the wall is
quite high, it may require a thick wall slab, for
example, or more. In this case a haunch
may be adopted, as shown below. This solution will
reduce the thickness of the wall, because it will be
designed for the moment at the section exactly
above the haunch.
The basement slab may have a thickness greater
than the wall thickness and may be extended
outside the wall by about or more, as
required.

13.5 LATERAL LOADS.

13.5.1 Introduction.

Loading on tall buildings differs from loading on low-rise buildings its accumulation into
much larger structural forces, in the increased significance of wind loading, and in the
greater importance of dynamic effects. The collection of gravity loading over a large number
of stories in a tall building can produce column loads of an order higher than those in low-
rise buildings. Wind loading on a tall building acts not only over a very large building surface,
but also with greater intensity at the greater heights and with a larger moment arm about
the base than on a low-rise building. Although wind loading on a low-rise building usually has
an insignificant influence on the design of the structure, wind on a high-rise building can
have a dominant influence on its structural arrangement and design. In an extreme case of a
528
Reinforced Concrete III Dr. Nasr Abboushi

very slender or flexible structure, the motion of the building in the wind may have to be
considered in assessing the loading applied by the wind.
In earthquake regions, any internal loads from the shaking of the ground may well exceed
the loading due to wind and, therefore, be dominant in influencing the building's structural
form, design, and cost. As an inertial problem, the building’s dynamic response plays a large
part in influencing, and in estimating, the effective loading on the structure.
With the exception of dead loading, the loads on a building cannot be assessed accurately.
While maximum gravity live loads can be anticipated approximately from previous field
observations, wind and earthquake loadings are random in nature, more difficult to measure
from past events, and even more difficult to predict with confidence. The application of
probabilistic theory has helped to rationalize, if not in every case to simplify, the approaches
to estimating wind and earthquake loading.

13.5.2 Wind Load.

The lateral loading due to wind or earthquake is the major factor that causes the design of
high-rise buildings to differ from those of low- to medium-rise buildings. For buildings of up
to about 10 stories and of typical proportions, the design is rarely affected by wind loads.
Above this height, however, the increase in size of the structural members, and the possible
rearrangement of the structure to account for wind loading, incurs a cost premium that
increases progressively with height. With innovations in architectural treatment, increases in
the strengths of materials, and advances in methods of analysis, tall building structures have
become more efficient and lighter and, consequently, more prone to deflect and even to
sway under wind loading. This served as a spur to research, which has produced significant
advances in understanding the nature of wind loading and in developing methods for its
estimation. These developments have been mainly in experimental and theoretical
techniques for determining the increase in wind loading due to gusting and the dynamic
interaction of structures with gust forces.

Response of Concrete Buildings to Wind Forces. The next Figure shows the path of wind
forces as they propagate through a simple, idealized concrete building. The windward wall
receives the wind pressures and transfers the resulting forces to the roof and floor
diaphragms. The diaphragms, in turn, transfer these loads to the elements of the lateral-
force-resisting system, which in this case, are the shear walls parallel to the wind forces. The
shear walls then transfer these loads to the building foundation. A similar scenario would
occur for frame buildings. In any case, the design of structural members for wind forces is
based on linear behavior; the structure is assumed to remain elastic under the design wind
forces.

529
Reinforced Concrete III Dr. Nasr Abboushi

13.5.3 Code provisions for wind loads.

In recent years, wind loads specified in codes and standards have been refined significantly.
This is because our knowledge of how wind affects buildings and structures has expanded
due to new technology and advanced research that have ensued in greater accuracy in
predicting wind loads. We now have an opportunity to design buildings that will satisfy
anticipated loads without excessive conservatism. The resulting complexity in the
determination of wind loads may be appreciated by comparing the 1973 Standard Building
Code (SBC), which contained only a page and one-half of wind load requirements, to the
2002 edition of the ASCE 7, which contains 97 pages of text, commentary, figures, and tables
to predict wind loads for a particular structure. As compared to a single method given in the
1973 SBC, ASCE 7 contains three methods for determining winds: the simplified procedure,
the analytical procedure, and the wind-tunnel procedure. The controlling equations for
determining wind loads require calculating velocity pressure as before, but are now modified
530
Reinforced Concrete III Dr. Nasr Abboushi

to account for several variables such as gusts, internal pressure, and aerodynamic properties
of the element under consideration, as well as topographic effects. Using the low-rise
buildings’ analytical procedure in ASCE-7 and applying it to the simplest building requires the
use of up to 11 variables. An important criterion that influences the calculation of wind loads
is the enclosure classification of the building. Three classifications are used: 1) enclosed;
2) partially enclosed; or 3) open. A building classified as partially enclosed assumes that a
large opening is on one side of a building and no (or minimal) openings are on the other
walls. As openings on one wall reach a certain size with respect to openings on the other
walls, the building is classified as partially enclosed. Depending upon the wind’s direction,
this type of situation allows two conditions to develop: internal pressure or internal suction.
Internal pressure occurs when air enters a building opening on the windward wall and
becomes trapped, exerting an additional force on the interior elements of the building.
Typically the internal pressures act in the same directions as the external pressures on all
walls except the windward wall. Internal suction is a condition that exists when there is an
opening on the leeward wall allowing air to be pulled out of the building. This results in the
internal forces acting in the same direction as the external forces on the windward wall. The
additional forces produced by this type of pressurization are characterized by requiring an
internal pressure coefficient that is more than three times greater than that required for an
enclosed building. Another criterion that significantly affects the magnitude of the wind
pressures is the site’s exposure category, which provides a way to define the relative
roughness of the boundary layers at the site.

The ASCE 7-05 and IBC-03 define three exposure categories: B, C, and D. Exposure B is the
roughest and D is the smoothest. Consequently, when all other conditions are equal,
calculated wind loads are reduced as the exposure category moves from D to B. Exposure B
is the most common category, consisting primarily of terrain associated with a suburban or
urban site. Accordingly, B is the default exposure category in both ASCE 7 and IBC. Exposure
C consists primarily of open terrain with scattered obstructions but also includes shoreline in
hurricane-prone regions. Exposure D applies to shore lines (excluding those in hurricane-
prone regions) with wind flowing over open water for a distance of at least one mile.
Buildings must also be classified based on their importance. The wind importance factor
specified in the codes is used to adjust the return period for a structure based on its relative
level of importance. For example, the importance factor for structures housing critical
national defense functions is , while the importance factor for an agricultural building
not as critical as a defense facility, is .
The applicable wind speeds for the United States and some tropical islands specified in the
wind speed maps are three-second gusts at ( ) above ground for Exposure
Category C. In the model codes that preceded the IBC (the National Building Code. Standard
Building Code, and Uniform Building Code) and versions of ASCE 7 prior to 1995, wind
speeds were shown as “fastest-mile winds,” which is defined as the average speed of a one-
mile column of air passing a reference point.

531
Reinforced Concrete III Dr. Nasr Abboushi

While the designated 3-sec gust wind speed for a particular site is higher than values on the
fastest-mile map, the averaging times are also different. The averaging time for a fastest-
mile wind speed is different for each wind speed, while the averaging time for the 3-sec gust
speeds varies from to , depending upon the sensitivity of the instruments.

At heights of approximately ( ) aboveground, the wind speed is virtually unaffected by surface


friction, and its movement is solely dependent on prevailing seasonal and local wind effects. The height
through which the wind speed is affected by topography is called the atmospheric boundary layer.

 Uniform Building Code, 1997: Wind Load Provisions

Wind load provisions of UBC 1997 are based on the ASCE 7-88 standard with certain
simplifying assumptions to make calculations easier. The design wind speed is based on the
fastest-mile wind speed as compared to the 3-sec gust speeds of the later codes. The
prevailing wind direction at the site is not considered in calculating wind forces on the
structures: The direction that has the most critical exposure controls the design.
Consideration of shielding by adjacent buildings is not permitted because studies have
532
Reinforced Concrete III Dr. Nasr Abboushi

shown that in certain configurations, the nearby buildings can actually increase the wind
speed through funneling effects or increased turbulence. Additionally, it is possible that
adjacent existing buildings may be removed during the life of the building being designed.
To shorten the calculation procedure, certain simplifying assumptions are made. These
assumptions do not allow determination of wind loads for flexible buildings that may be
sensitive to dynamic effects and wind-excited oscillations such as vortex shedding. Such
buildings typically are those with a height-to-width ratio greater than 5, and over 400 ft
( ) in height. The general section of the UBC directs the user to an approved standard
for the design of these types of structures. The ASCE 7-05, adopted by IBC 2003, is one such
standard for determining the dynamic gust response factor required for the design of these
types of buildings. UBC provisions are not applicable to buildings taller than ( )
for normal force method, Method 1, and ( ) for projected area method, Method
2. Any building, including those not covered by the UBC, may be designed using wind-tunnel
test results.

 Wind Speed Map

The minimum basic wind speed at any site in the country can be defined from the wind
speed map. The wind speed represents the fastest-mile wind speed in an exposure C terrain
at ( ) above grade, for a 50-year mean recurrence interval. The probability of
experiencing a wind speed faster than the value indicted in the map, in any given year is
1 in 50, or .

 Special Wind Regions

Although basic wind speeds are constant over hundreds of miles, some areas have local
weather or topographic characteristics that affect design wind speeds. These special wind
regions are defined in the UBC map. Because some jurisdictions prescribe basic wind speeds
higher than the map, it is prudent to contact local building officials before commencing with
the wind design.

 Exposure Effects

Every building site has its own unique characteristics in terms of surface roughness and
length of upwind terrain associated with the roughness. Simplified code methods cannot
account for the uniqueness of the site. Therefore the code approach is to assign broad
exposure categories for design purposes.
Similar to the ASCE method, the UBC distinguishes between three exposure categories; B, C,
and D. Exposure B is the least severe, representing urban, suburban, wooded, and other
533
Reinforced Concrete III Dr. Nasr Abboushi

terrain with numerous closely spaced surface irregularities; Exposure C is for flat and
generally open terrain with scattered obstructions; and the most severe, Exposure D, is four
unobstructed coastal areas directly exposed to large bodies of water. Discussion of the
exposure categories follows.
It should be noted that Exposure A (centers of large cities where over half the buildings have
a height in excess of ( )), included in some standards, is not recognized in the
UBC. The UBC considers this type of terrain as Exposure B, allowing no further decrease in
wind pressure.
Exposure B has terrain with buildings, forest, or surface irregularities, covering at least
of the ground level area extending ( ) or more from the site.
Exposure C has terrain that is flat and generally open, extending one-half mile (0.81 km) or
more from the site in any full quadrant.
Exposure D represents the most severe exposure in areas of basic wind speeds of
( ) or greater, and has terrain that is flat and unobstructed facing large bodies of
water over one mile ( ) or more in width relative to any quadrant of the building
site. Exposure D extends inland from shoreline one-fourth mile ( ) or 10 times the
building height, whichever is greater.

 Site Exposure

Even though a building site may have different exposure categories in different directions,
the most severe exposure is used for all wind-load calculations regardless of building
orientation or direction of wind.
Exposure D is perhaps the easiest to determine because it is explicitly for unobstructed
coastal areas directly exposed to large bodies of water. It is not as easy to determine
whether a site falls into Exposure B or C because the description of these categories is
somewhat ambiguous. Moreover, the terrain surrounding a site is usually not uniform and
can be composed of zones that would be classified as Exposure B while others would be
classified as Exposure C. When such a mix is encountered, the more severe exposure
governs. The UBC classifies a site as Exposure C when open terrain exists for one full
quadrant extending outward from the building for at least one-half mile. If the quadrant is
less than or less than one-half mile, then the site is classified as Exposure B. It is
essential to select the appropriate category because force levels could differ by as much as
65% between Exposure B and C. It is advisable to contact the local building official before
embarking on a building design with a questionable site exposure category. If the site has a
view of a cliff or hill, it may be prudent to assign Exposure C to D to account for higher wind
velocity effects.

534
Reinforced Concrete III Dr. Nasr Abboushi

 Design Wind Pressures

The design wind pressure is given as a product of the combined height, exposure, and gust
factor coefficient ; the pressure coefficient ; the wind stagnation pressure ; and
building Importance Factor .

The pressure manifesting on the surface of a building due to a mass of air with density ,
moving at a velocity is given by Bernoulli’s equation

The density of air is ( ), for conditions of standard atmosphere,


temperature ( ), and barometric pressure ( ).
Since velocity given in the wind map is in , the above equation for reduces to

[ ][ ] ( )

( )

For instance, if the wind speed is , , which the


UBC rounds off to (see next table). Note UBC does not consider the effect of
reduced air density at sites located at higher altitudes.

 The Factor

The effects of height, exposure, and gust factor are all lumped into one factor in the
interest of keeping the UBC method simple. Values of shown in Table below (UBC, Table
16-G) are essentially equal to the product of two parameters , the velocity pressure
exposure coefficient, and , the gust response factor. Both these parameters are defined
separately in ASCE 7-05, and hence are more appropriate for non-ordinary buildings.
The height and exposure factors account for the terrain effects on gradient heights and
typically cause lower wind speeds in built-up terrain than in an open terrain. The gust factor
accounts for air turbulence and dynamic building behavior.
For low-rise buildings with natural period of less than 1 sec, the wind response is essentially
static with the lateral deflection proportional to the wind force. For tall buildings, on the
other hand, the response is dynamic resulting in deflections greater than those estimated by
simple procedures. Therefore for slender buildings a procedure such as the one given in the
535
Reinforced Concrete III Dr. Nasr Abboushi

ASCE 7-05, which takes into account the dynamic characteristic of the building, would likely
to be more appropriate

Combined height, exposure and gust factor coefficient ( )

Height above average level of


Adjoining ground
Exposure D Exposure C Exposure B
[feet] [meter]

1
Values for intermediate heights above ( ) may be interpolated

 Pressure Coefficient

The given in Table below is a function of building shape and location, and whether the
wind load induces inward or outward pressures.
It is given in two parts. The first part, Primary Frames Systems, is for the design of the entire
building. The second part, Elements and Components of Structure, is for the design of
cladding.
Wind gusting around a building does not cause peak pressures and sections simultaneously
over the entire surface of the building. Therefore, wind loads for design of primary frames
and systems are calculated using average wind pressures and suctions. On the other hand
the design of building components such as curtain walls and cladding is controlled by the
instantaneous peak pressures and suction acting over relatively small localized areas. This is
the reason why the pressures and suctions for building components are larger than those for
the entire building.
536
Reinforced Concrete III Dr. Nasr Abboushi

537
Reinforced Concrete III Dr. Nasr Abboushi

𝐶𝑞 (𝑢𝑝𝑤𝑎𝑟𝑑)

𝐶𝑞 (𝑖𝑛𝑤𝑎𝑟𝑑)

𝐶𝑞 (𝑜𝑢𝑡𝑤𝑎𝑟𝑑)
Wind Direction

Elevation

Wind pressures and suctions for primary systems are mainly a function of the building
height. Although these are influenced by the building’s shape, the roughness of its exterior,
and its plan aspect ratio, these are ignored. For example, even though wind load on a
circular building is theoretically about 80% of that for a rectangular building, no reduction of
forces is permitted in the UBC.
Two methods, are given in the UBC for determining wind loads for primary frames. Method
1, the normal force method, is applicable to all structures, and is the only method permitted
for the design of gable-roofed buildings. It assumes wind loads act perpendicular to the
surfaces of the roof, and the walls. Method 2, the projected area method, is easier to use
than Method 1. The wind pressures and suctions are integrated into a single value and are
assumed to act on the entire projected area of the building, instead of on individual surfaces
of roof and walls.
Another important difference between the two methods is that method 1 uses a constant
value of based on mean roof height to calculate wind suctions on leeward walls.
Method 2 uses a value that varies with height. Hence, method 2 underestimates the wind
loads on taller structures. For this reason, use of method 2 is limited to structures less than
( ), in order to minimize the underestimated leeward forces.
538
Reinforced Concrete III Dr. Nasr Abboushi

 Importance Factor
Importance factor is applied to increase the wind loads for certain occupancy categories.
The 1997 UBC gives five separate occupancy categories: essential facilities, hazardous
facilities, special occupancy structures, standard occupancy structures, and miscellaneous
structures. Essential or hazardous facilities are assigned an importance factor ,
which has the effect of increasing the mean reference interval from a -year to a -year
return period. Special structures, standard occupancy structures, and miscellaneous
structures are assigned an importance factor of . Office and residential buildings are
typically assigned a standard occupancy factor of .

Example.
Compare the design wind pressures of UBC methods 1 and 2 for the primary lateral force
resisting system for the building in figure. This is a gable structure that does not have a
system of rigid frames for resisting lateral forces. The building is a standard occupancy
enclosed structure.
Exposure category , Basic wind speed
Note:

solution:

• Wind stagnation pressure UBC Table 16-F (see


table on page 535)
• Importance factor UBC Table 16-K
𝑚 ( 𝑓𝑡)
𝑓𝑡)
𝑚(

𝑚( 𝑓𝑡)

• Total height of building


• Combined height, exposure, and gust coefficient

539
Reinforced Concrete III Dr. Nasr Abboushi

• Pressure coefficients are obtained from UBC Table 16-H (see page 537)

Method 1 (Normal Force method):

 Windward wall

 Leeward wall
For Leeward wall is a constant and based on the mean height of the roof

For

 Windward roof

or .

 Leeward roof

Method 2 (Projected Area method):

Here we use instead of doing interpolation for .

Uplift pressure upward

540
Reinforced Concrete III Dr. Nasr Abboushi

𝑝𝑠𝑓

𝑝𝑠𝑓
𝑝𝑠𝑓

Method 1
𝑝𝑠𝑓 Normal Force method 𝑝𝑠𝑓

Wind direction
𝑝𝑠𝑓

𝑝𝑠𝑓

Method 2
𝑓𝑡

Projected Area method 𝑝𝑠𝑓

Example (Eleven-Story Building: UBC 1997).


Given.
 Eleven-story communication building deemed necessary for post-disaster emergency
communications,
 Building height ( ) consisting of 2 bottom floors at ( ) and 9
typical floors at ( )
 Exposure category
 Basic wind speed
 Building width
541
Reinforced Concrete III Dr. Nasr Abboushi

Required. Design wind pressures on primary wind-resisting system.

solution:
The design pressure is given by the chain equation

The values of the combined height, exposure, and gust factor coefficient tabulated in
Table below are taken directly from Table on page 128. Note that for suction on the leeward
face, is at the roof hight, and is constant for the full height of the building. The wind
pressure corresponding to basic wind speed of 100 mph is given by

The values of pressure coefficient , obtained using the normal force method (Method 1),
are 0.8 for the inward pressure on the windward face, and 0.5 for the suction on the leeward
face. Because the building is less than ( ), the combined value of
may be used throughout the height to calculate the wind load on the
primary wind-resisting system. Observe that Method 2 (projected area method) yields the
same value of
Design pressures and floor-by-floor wind loads are shown in Table below. Notice that the
wind pressure and suction on the lower half of the first story (between the ground and
aboveground) is commonly considered to be transmitted directly into the ground.

The wind load at each level is obtained by multiplying the tributary area for the level by the
average of design pressures above and below that level. For example,

( )

542
Reinforced Concrete III Dr. Nasr Abboushi

𝑝𝑠𝑓
𝑝𝑠𝑓
𝑝𝑠𝑓
𝑝𝑠𝑓
𝑝𝑠𝑓
𝑝𝑠𝑓 𝑝𝑠𝑓

𝑝𝑠𝑓
𝑝𝑠𝑓
𝑝𝑠𝑓
𝑝𝑠𝑓

𝑝𝑠𝑓

 ASCE 7-05: Wind Load Provisions

The full title of this ASCE standard is American Society of Civil Engineers Minimum Design
Loads for Buildings and Other Structures. In one of its 10 sections, ASCE 7-05 provides three
procedures for calculating wind loads for buildings and other structures, including the main
wind-force-resisting systems and all components thereof. The designer can use Method 1,
the simplified procedure, to select wind pressures directly without calculation when the
building is less than in height and meets all requirements given in Section 6.4 of the
standard. Method 2 can be used for buildings and structures of any height that are regular in
shape, provided the buildings are not sensitive to across-wind loading, vortex shedding, or
instability due to galloping or flutter; or do not have a site for which channeling effects
warrant special consideration. Method 3 is a wind-tunnel test procedure that can be used in
lieu of methods 1 and 2 for any building or structure. Method 3 is recommended for
buildings that possess any of the following characteristics:
 Have nonuniform shapes.
 Are flexible with natural frequencies less than .
 Are subject to significant buffeting by the wake of upwind buildings or other
structures.
 Are subject to accelerated flow of wind by channeling or local topographic features.
Basic wind speeds for are shown on a map having isotachs representing a 3-sec gust speed at
( ) above the ground. The map is standardized to represent a 50-year recurrence
interval for exposure C topography (flat, open, country and grasslands with open terrain and
scattered obstructions generally less than ( ) in height). The minimum basic wind

543
Reinforced Concrete III Dr. Nasr Abboushi

speed provided in the standard is ( ). Increasing the minimum wind speed for
special topographies such as mountain terrain, gorges, and ocean fronts is recommended.
The wind speed map for the United States and adjoining landmasses is based on data
collected over a long period of time at weather stations located throughout the country. The
maximum wind velocity expected at any location can be found simply by referring to the
map.
The abandonment of the fastest-mile speed in favor of a 3-sec-gust speed first took place in
the ASCE 7-1995 edition. The reasons are: 1) modern weather stations no longer measure
wind speeds using the fastest-mile method; 2) a 3-sec-gust speed is closer to the sensational
wind speeds often quoted by news media; and 3) it matches closely the wind speeds
experienced by small buildings and by components of all buildings.
Method 1, the simplified procedure, and Method 3, the wind tunnel procedure, are not
discussed here.
Method 2, the analytical procedure which is not covered in this section, applies to a majority
of buildings. It accounts for the following factors that influence the design wind forces:
1. The basic wind speed.
2. The mean recurrence interval of the wind speed considered appropriate for the
design.
3. The characteristics of the terrain surrounding the building.
4. The height at which the wind load is being determined.
5. Directional properties of the wind climate.
6. The size, geometry, and aerodynamics of the building.
7. The positions of the area acted on by the wind flow.
8. The magnitude of the area of interest.
9. The porosity of the building envelope.
10. The structural properties that may make the building susceptible to dynamic effects.
11. The speed-up effect of certain topographic features such as hills and escarpments.

The analytical procedure has two steps. The first step considers the properties of the wind
flow and the second accounts for the properties of the structure and its dynamic response to
the longitudinal (along-wind) wind turbulence. The effects of across-wind response are not
explicitly considered in the ASCE 7-05, Methods 1 and 2. For details see the ASCE 7-05
standard.

Wind-tunnel testing of buildings has been an offshoot of aeronautical engineering, in which


the flow of wind is duplicated at high altitudes. The tunnels for testing airplanes are
designed to minimize the effects of turbulence, and as such, they do not duplicate
atmospheric boundary layer or wind turbulence. This is because majority of airplane flights,
except for brief periods of landing and takeoff, occur at a height well above the boundary
layer. Building activity, on the other hand, occurs precisely within this atmospheric boundary
layer, characterized by a gradual retardation of wind speed and high turbulence near the
surface of the earth. Therefore, for testing of buildings, aeronautical wind tunnels have been
544
Reinforced Concrete III Dr. Nasr Abboushi

modified and entirely new facilities have been built to reproduce turbulence and natural
flow of wind within the boundary layer.
Wind-tunnel tests (or similar tests employing fluids other than air) are considered to be
properly conducted only if the following conditions are satisfied:
1. The natural atmospheric boundary layer has been modeled to account for the
variation of wind speed with height.
2. The length scale of the longitudinal component of atmospheric turbulence is
modeled to approximately the same scale as that used to model the building.
3. The modeled building and surrounding structures and topography are geometrically
similar to their full-scale counterparts.
4. The projected area of the modeled building and surroundings is less than 8% of the
test section cross-sectional area unless correction is made for blockage.
5. The longitudinal pressure gradient in the wind tunnel test section is accounted for.
6. Reynolds number effects on pressures and forces are minimized.
7. Response characteristics of the wind-tunnel instrumentation are consistent with the
required measurements.

Boundary-layer wind tunnels capable of developing flows that meet the conditions
stipulated above typically have test-section dimensions in the following ranges: width,
( ); height, ( ); and length, ( ).
Maximum wind speeds are ordinarily in the range of ( ).
Three basic types of wind-tunnel test models are commonly used:
1. Rigid pressure model (PM)
2. Rigid high-frequency base balance model (H-FBBM)
3. Aeroelastic model (AM)

One or more of the models may be employed to obtain design loads for a particular building
or structure. The pressure model provides local peak pressures for design of elements such
as cladding and mean pressures for the determination of overall mean loads.
The high-frequency model measures overall fluctuating loads for the determination of
dynamic responses. The aeroelastic model is employed for direct measurement of overall
loads, deflections, and accelerations, when the lateral motions of a building are considered
to have a large influence on wind loading.
Various techniques are used in aeronautical tunnels to generate turbulence and atmospheric
boundary layer by using devices such as screens, spires, and grids. In special wind tunnels
with long test sections, turbulent boundary layer is generated by installing appropriate
roughness elements in the upstream flow. Another approach is to use a counterjet
technique. In every case there is some question whether the natural wind turbulence
characteristic is appropriately modeled and proper gust simulation is included. The degree of
scaling required to appropriately account for these may yield a very extreme scale for the
building, on the order of 1:500 or even more for urban environment studies.

545
Reinforced Concrete III Dr. Nasr Abboushi

546
Reinforced Concrete III Dr. Nasr Abboushi

547

You might also like