0% found this document useful (0 votes)
50 views83 pages

Appendices Seismic Guide 100518

This document provides background on seismic response of structures to earthquakes. It discusses elastic response and concepts like natural frequency, period, damping ratio and response spectra. Response spectra plot the maximum displacement or acceleration that structures with different periods will experience when subjected to earthquake ground motion. The document includes figures illustrating development of displacement response spectra and examples of displacement and acceleration response spectra for the 1940 El Centro earthquake. It also shows the uniform hazard response spectrum used in the National Building Code of Canada 2005.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
50 views83 pages

Appendices Seismic Guide 100518

This document provides background on seismic response of structures to earthquakes. It discusses elastic response and concepts like natural frequency, period, damping ratio and response spectra. Response spectra plot the maximum displacement or acceleration that structures with different periods will experience when subjected to earthquake ground motion. The document includes figures illustrating development of displacement response spectra and examples of displacement and acceleration response spectra for the 1940 El Centro earthquake. It also shows the uniform hazard response spectrum used in the National Building Code of Canada 2005.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 83

SEISMIC DESIGN GUIDE FOR MASONRY BUILDINGS

Second Edition

Svetlana Brzev Donald Anderson

Canadian Concrete Masonry Producers Association

2018
TABLE OF CONTENTS – APPENDIX A

A. RESPONSE OF STRUCTURES TO EARTHQUAKES ................................................ A-2

A.1. Elastic Response ........................................................................................................................ A-2

A.2. Inelastic Response ...................................................................................................................... A-6

A.3. Ductility ........................................................................................................................................ A-7

A.4. A Primer on Modal Dynamic Analysis Procedure.................................................................... A-8


A.4.1. Multi-degree-of-freedom systems ......................................................................................... A-8
A.4.2. Seismic analysis methods ..................................................................................................... A-9
A.4.3. Modal analysis procedure: an example .............................................................................. A-10
A.4.4. Comparison of static and modal analysis results ................................................................ A-14

9/1/2018 A-1
A. Response of Structures to Earthquakes
This appendix contains background related to fundamentals of seismic response of structures to
earthquakes. A discussion on elastic and inelastic response is included, and a primer on modal
dynamic analysis.

A.1. Elastic Response


When an earthquake strikes, the base of a building is subject to lateral motion while the upper
part of the structure initially is at rest. The forces created in the structure from the relative
displacement between the base and upper portion cause the upper portion to accelerate and
displace. At each floor the lateral force required to accelerate the floor mass is provided by the
forces in the vertical members. The floor forces are inertial forces, not externally applied forces
such as wind loads, and exist only as long as there is movement in the structure.

Earthquakes cause the ground to shake for a relatively short time, 15 to 30 seconds of strong
ground shaking, although large subduction earthquakes may last for a few minutes. The motion
is cyclic and the response of the structure can only be determined by considering the dynamics
of the problem. A few important dynamic concepts are discussed below.

Consider a simple single-storey building with masonry walls and a flat roof. The building can be
represented by a Single Degree of Freedom (SDOF) model (also known as a stick model) as
shown in Figure A-1a). The mass, M , lumped at the top, represents the mass of the roof and a
fraction of the total wall mass, while the column represents the combined wall stiffness, K , in
the direction of earthquake ground motion. If an earthquake causes a lateral deflection,  , at
the top of the building, Figure A-1b), and if the building response is elastic with stiffness, K ,
then the lateral inertial force, F , acting on the mass M will be

F  K 

When the mass of a SDOF un-damped structure is allowed to oscillate freely, the time for a
structure to complete one full cycle of oscillation is called the period, T , which for the SDOF
system shown is given by
M
T  2 (seconds)
K

Instead of period, the term natural frequency,  , is often used in seismic design. It is related to
the period as follows

2 K
  (radians/sec)
T M

Frequency is sometimes also expressed in Hertz, or cycles per second, instead of radians/sec,
denoted by the symbol  cps , where

1 
cps  
T 2

9/1/2018 A-2
Figure A-1. SDOF system: a) stick model; b) displaced position.
As the structure vibrates, there is always some energy loss which will cause a decrease in the
amplitude of the motion over time - this phenomenon is called damping. The extent of damping
in a building depends on the materials of construction, its structural system and detailing, and
the presence of architectural components such as partitions, ceilings and exterior walls.
Damping is usually modelled as viscous damping in elastic structures, and hysteretic damping
in structures that demonstrate inelastic response. In seismic design of buildings, damping is
usually expressed in terms of a damping ratio,  , which is described in terms of a percentage of
critical viscous damping. Critical viscous damping is defined as the level of damping which
brings a displaced system to rest in a minimum time without oscillation. Damping less than
critical, an under-damped system, allows the system to oscillate; while an over-damped system
will not oscillate but take longer than the critically damped system to come to rest. Damping has
an influence on the period of vibration, T, however this influence is minimal for lightly damped
systems, and in most cases, is ignored for structural systems. For building applications, the
damping ratio can be as low as 2%, although 5% is used in most seismic calculations where
some nonlinear response is present. Damping in a structure increases with displacement
amplitude since with increasing displacement more elements may crack or become slightly
nonlinear. For linear seismic analysis viscous damping is usually taken as 5% of critical as the
structural response to earthquakes is usually close to or greater than the yield displacement. A
smaller value of viscous damping is usually used in non-linear analyses as hysteretic damping
is also considered.

One of the most useful seismic design concepts is that of the response spectrum. When a
structure, say the SDOF model shown in Figure A-1, is subjected to an earthquake ground
motion, it cycles back and forth. At some point in time the displacement relative to the ground
and the absolute acceleration of the mass reach a maximum,  max and a max , respectively.
Figure A-2a) shows the maximum displacement plotted against the period, T . Denote the
period of this structure as T1 . If the dynamic properties, i.e. either the mass or stiffness change,
the period will change, say to T2 . As a result, the maximum displacement will change when the
structure is subjected to the same earthquake ground motion, as indicated in Figure A-2b).
Repeating the above process for many different period values and then connecting the points
produces a plot like that shown in Figure A-2c), which is termed the displacement response
spectrum. The spectrum so determined corresponds to a specific input earthquake motion and a
specific damping ratio,  . The same type of plot could be constructed for the maximum
acceleration, a max , rather than the displacement, and would be termed the acceleration
response spectrum.

9/1/2018 A-3
Figure A-2. Development of a displacement response spectrum - maximum displacement
response for different periods T : a) T  T1 ; b) T  T2 ; c) many values of T .
Figure A-3a) shows the displacement response spectrum for the 1940 El Centro earthquake at
different damping levels. Note that the displacements decrease with an increase in the damping
ratio,  , from 2% to 10%. Figure A-3b) shows the acceleration response spectrum for the same
earthquake. For the small amount of damping present in the structures, the maximum
acceleration, a max , occurs at about the same time as the maximum displacement,  max , and
these two parameters can be related as follows
2
 2 
a max    max
 T 
Thus, by knowing the spectral acceleration, it is possible to calculate the displacement spectral
values and vice versa. It is also possible to generate a response spectrum for maximum
velocity. Except for very short and very long periods, the velocity, v max , is closely approximated
by
 2 
v max     max
T 
This is generally called the pseudo velocity response spectrum as it is not the true velocity
response spectrum.

9/1/2018 A-4
a)

b)
Figure A-3. Response spectra for the 1940 El Centro NS earthquake at different damping levels:
a) displacement response spectrum; b) acceleration response spectrum.
The response spectrum can be used to determine the maximum response of a SDOF structure,
given its fundamental period and damping, to a specific earthquake acceleration record.
Different earthquakes produce widely different spectra and so it has been the practice to choose
several earthquakes (usually scaled) and use the resulting average response spectrum as the
design spectrum. For years, the NBC seismic provisions have used this procedure where the
design spectrum for a site was described by one or two parameters, either peak ground
acceleration and/or peak ground velocity, that were determined using probabilistic means.

9/1/2018 A-5
More recently, probabilistic methods have been used to determine the spectral values at a site
for different structural periods. Figure A-4 shows the 5% damped acceleration response
spectrum for Vancouver used in developing the NBC 2005. This is a uniform hazard response
spectrum, i.e., spectral accelerations corresponding to different periods are based on the same
probability of being exceeded, that is, 2% in 50 years. This is discussed further in Section 1.3.
The NBC 2015 code uses the same method but has been updated by using many more records
to determine the hazard and has extended the period range out to 10 seconds.

Figure A-4. Uniform hazard acceleration response spectrum for Vancouver, 2% in 50 year
probability, 5% damping.

A.2. Inelastic Response


For any given earthquake ground motion and SDOF elastic system it is possible to determine
the maximum acceleration and the related inertia force, Fel , and the maximum displacement,
 el . Figure A-5a) shows a force-displacement relationship with the maximum elastic force and
displacement indicated. If the structure does not have sufficient strength to resist the elastic
force, Fel , then it will yield at some lower level of inertia force, say at lateral force level, Fy . It
has been observed in many studies that a structure with a nonlinear cyclic force-displacement
response similar to that shown in Figure A-5b) will have a maximum displacement that is not
much different from the maximum elastic displacement. This is indicated in Figure A-5c) where
the inelastic (plastic) displacement,  u , is shown just slightly greater than the elastic
displacement,  el . This observation has led to the equal displacement rule, an empirical rule
which states that the maximum displacement that the structure reaches in an earthquake is
independent of its yield strength, i.e. irrespective of whether it demonstrates elastic or inelastic
response. The equal displacement rule is thought to hold because the nonlinear response
softens the structure and so the period increases, thereby giving rise to increased
displacements. However, at the same time, the yielding material dissipates energy that
effectively increases the damping (the energy dissipation is proportional to the area enclosed by
the force-displacement loops, termed hysteresis loops). Increased damping tends to decrease
the displacements; therefore, it is possible that the two effects balance one another with the
result that the elastic and inelastic displacements are not significantly different.

9/1/2018 A-6
Figure A-5. Force-displacement relationship: a) elastic response; b) nonlinear (inelastic)
response; c) equal displacement rule.
There are limits beyond which the equal displacement rule does not hold. In short period
structures, the nonlinear displacements are greater than the elastic displacements, and for very
long period structures, the maximum displacement is equal to the ground displacement.
However, the equal displacement rule is, in many ways, the basis for the seismic provisions in
many building codes which allow the structure to be designed for forces less than the elastic
forces. But there is always a trade-off, and the lower the yield strength, the larger the nonlinear
or inelastic deformations. This can be inferred from Figure A-5c) where it is noted that the
difference between the nonlinear displacement,  u , and yield displacement,  y , which
represents the inelastic deformation, would increase as the yield strength decreases. Inelastic
deformations generally relate to increased damage, and the designer needs to ensure that the
strength does not deteriorate too rapidly with subsequent loading cycles, and that a brittle failure
is prevented. This can be achieved by additional “seismic” detailing of the structural members,
which is usually prescribed by the material standards. For example, in reinforced concrete
structures, seismic detailing consists of additional confinement reinforcement that ensures
ductile performance at critical locations in beams, columns, and shear walls. In reinforced
masonry structures, it is difficult to provide similar confinement detailing, and so restrictions are
placed on limiting the reinforcement spacing, on levels of grouting, and on certain strain limits in
the masonry structural components (e.g. shear walls) which provide resistance to seismic loads
(see Chapter 2 for more details on seismic design of masonry shear walls).

A.3. Ductility
Ductility relates to the capacity of the structure to undergo inelastic displacements. For the
SDOF structure, whose force-displacement relation is shown in Figure A-5c) the displacement
ductility ratio,   , is a measure of damage that the structure might undergo and can be
expressed as

u
 
y

The ratio between the maximum elastic force, Fel , and the yield force, Fy , is given by the force
reduction factor, R , defined as

9/1/2018 A-7
Fel
R
Fy

If the material is elastic-perfectly plastic, i.e. there is no strain hardening as it yields (see Figure
A-5b), and if  u is equal to  el , then it can be shown that   is equal to R .

For different types of structures and detailing requirements, most building codes tend to
prescribe the R value while not making reference to the displacement ductility ratio,   , thus
implying that the   and R values would be similar.

A.4. A Primer on Modal Dynamic Analysis Procedure


The main objective of this section is to explain how more complex multi-degree-of-freedom
structures respond to earthquake ground motions and how such response can be quantified in a
form useful for structural design. This background should be helpful in understanding the NBC
seismic provisions.

A.4.1. Multi-degree-of-freedom systems


The idea of modelling the building as a SDOF structure was introduced in Section A.1, and the
dynamic response to earthquake ground motions was developed in terms of a response
spectrum. Such a simple model might well represent the lateral response of a single storey
warehouse building with flexible walls or bracing system, and with a rigid roof system where the
roof comprises most of the weight (mass) of the structure. However, this is not a good model for
a masonry warehouse with a metal deck roof, where the walls are quite stiff and the deck is
flexible and light relative to the walls. Such a system requires a more complex model using a
multi-degree-of-freedom (MDOF) system. A shear wall in a multi-storey building is another
example of a MDOF system.

Figure A-6 shows two examples of MDOF structures. A simple four-storey structure is shown in
Figure A-6a), and a simple MDOF model for this building consists of a column representing the
stiffness of vertical members (shear walls or frames), with the masses lumped at the floor levels.
If the floors are rigid, it can be assumed that the lateral displacements at every point in a floor
are equal, and the structure can be modelled with one degree of freedom (DOF) at each floor
level (a DOF can be defined as lateral displacement in the direction in which the structure is
being analyzed). This will result in as many degrees of freedom as the number of floors, so this
building can be modelled as a 4-DOF system. It must also be assumed that there are no
torsional effects, that is, there is no rotation of the floors about a vertical axis (torsional effects
are discussed in Section 1.11). The analysis will be the same irrespective of the lateral force
resisting system (a shear wall or a frame), aside from details in finding the lateral stiffness
matrix for the floor displacements.

The warehouse building shown in Figure A-6b) is another example of a MDOF structure. The
walls are treated as a single column with some portion of the wall and roof mass, M 1 , located at
the top. The roof can be treated as a spring (or several springs) with the remaining roof mass,
M 2 , attached to the spring(s). How much mass to attach to each degree of freedom, and how
to determine the stiffness of the roof, are major challenges in this case.

9/1/2018 A-8
Figure A-6. MDOF systems: a) multi-storey shear wall building; b) warehouse with flexible roof.

A.4.2. Seismic analysis methods


The question of interest to structural engineers is how to determine a realistic seismic response
for MDOF systems? The possible approaches are:
 static analysis, and
 dynamic analysis (modal analysis or time history method).

The simplest method is the equivalent static analysis procedure (also known as the quasi-static
method) in which a set of static horizontal forces is applied to the structure (similar to a wind
load). These forces are meant to emulate the maximum effects in a structure that a dynamic
analysis would predict. This procedure works well when applied to small, simple structures, and
also to larger structures if they are regular in their layout.

NBC 2015 specifies a dynamic analysis as the default method. The simplest type of dynamic
analysis is the modal analysis method. This method is restricted to linear systems, and consists
of a dynamic analysis to determine the mode shapes and periods of the structure, and then

9/1/2018 A-9
uses a response spectrum to determine the response in each mode. The response of each
mode is independent of the other modes, and the modal responses can then be combined to
determine the total structural response. In the next section, the modal analysis procedure will be
explained with an example.

The second type of dynamic analysis is the time history method. This consists of a dynamic
analysis model subjected to a time-history record of an earthquake ground motion. Time history
analysis is a powerful tool for analyzing complex structures and can take into account nonlinear
structural response. This procedure is complex and time-consuming to perform, and as such,
not warranted for low-rise and regular structures. It requires an advanced level of knowledge of
the dynamics of structures and it is beyond the scope of this document. For detailed background
on dynamic analysis methods the reader is referred to Chopra (2007).

A.4.3. Modal analysis procedure: an example


Consider a four-storey shear wall building example such as that shown in Figure A-6a). The
building can be modelled as a stick model, with a weight, W , of 2,000 kN lumped at each floor
level, and a uniform floor height of 3 m (see Figure A-7). For simplicity, the wall stiffness and the
masses are assumed uniform over the height. This model is a MDOF system with four degrees
of freedom consisting of a lateral displacement at each storey level. A MDOF system has as
many modes of vibration as degrees of freedom. Each mode has its own characteristic shape
and period of vibration. The periods are given in Table A-1, the four mode shapes are given in
Table A-2 and shown in Figure A-7. In this example, the stiffness has been adjusted to give a
first mode period of 0.4 seconds, which is representative of a four-storey structure based on a
simple rule-of-thumb that the fundamental period is on the order of 0.1 sec per floor. Note that
the first mode, also known as the fundamental mode, has the longest period. The first mode is
by far the most important for determining lateral displacements and interstorey drifts, but higher
modes can substantially contribute to the forces in structures with longer periods. In this
example the mode shapes have been normalized so that the largest modal amplitude is unity.

For linear elastic structures, the equations governing the response of each mode are
independent of the others provided that the damping is prescribed in a particular manner. Thus,
the response in each mode can be treated in a manner similar to a SDOF system, and this
allows the maximum displacement, moment and shear to be calculated for each mode. In the
final picture, the modal responses have to somehow be combined to find the design forces (this
will be discussed later in this section). Modal analysis can be performed by hand calculation for
a simple structure, however, in most cases, the use of a dynamic analysis computer program
would be required.

Knowing the mode shapes and the mass at each level, it is possible to calculate the modal
mass for each mode, which is given in Table A-1 as a fraction of the total mass of the structure.
The modal masses are representative of how the mass is distributed to each mode, and the
sum of the modal masses must add up to the total mass. When doing modal analysis, a
sufficient number of modes should be considered so that the sum of the modal masses adds up
to at least 90% of the total mass. In the example here this would indicate that only the first two
modes would need to be considered (0.696 + 0.210 = 0.906).

9/1/2018 A-10
Figure A-7. Four-storey shear wall building model and modal shapes.
As an example of how the different modes can be used to determine the structural response,
Figure A-8 shows a typical design acceleration response spectrum which will be used to
determine the modal displacements and accelerations. The four modal periods are indicated on
the spectrum (note that only the first two periods are identified on the diagram; T1=0.40 and
T2=0.062 sec) and the spectral acceleration Sa at each of the periods is given in Table A-3.

Figure A-8. Design acceleration response spectrum.


A very useful feature of the modal analysis procedure is that it gives the base shear in each
mode as a product of the modal mass and the spectral acceleration Sa for that mode, as shown
in Table A-3. For example, the base shear for the first mode is equal to (8000kN x 0.696) x 0.74
= 4127 kN). Note that the spectral acceleration is higher for the higher modes, but because the
modal mass for these modes is smaller, the base shear is smaller. The inertia forces from each
floor mass act in the same directions as the mode shape, that is, some forces are positive while
others are negative for the higher modes (refer to mode shapes shown in Figure A-7). It can be
seen from the figure that the forces from the first mode all act in the same direction at the same
time, while the higher modes will have both positive and negative forces. Thus, the base shear
from the first mode is usually larger than that from the other modes.

9/1/2018 A-11
The modal base shears shown in Table A-3 are the maximum base shears for each mode. It is
very unlikely that these forces will occur at the same time during the ground shaking, and they
could have either positive or negative signs. Summing the contribution of each mode where all
values are taken as positive, known as the absolute sum (ABS) method, produces a very high
upper bound estimate of the total base shear. Statistical analyses have shown that the square-
root-of-the-sum-of-the squares (RSS) procedure, where the contribution of each mode is
squared, and the square root is then taken of the sum of the squares, gives a reasonably good
estimate of the modal sum, especially if the modal periods are widely separated.

Table A-3 shows the base shear values estimated by the two methods and gives an indication
of the conservatism of the ABS method for this case (total base shear of 6,462 kN), where the
modal periods are widely separated, and use of the RSS method is appropriate since it gives a
lower total base shear value of 4,468 kN. Note that there is a third method that is incorporated in
many modal analysis programs called the complete-quadratic-combination (CQC) method. This
method should be used if the periods of some of the modes being combined are close together,
as would be the case in many three-dimensional structural analyses, but for most structures
with well-separated periods and low damping, the result of the RSS and CQC methods will be
nearly identical (this is true for most two-dimensional structural analyses).

The amplitude of displacement in each mode is dependent upon the spectral acceleration for
that mode and its modal participation factor, which is a measure of the degree to which a certain
mode participates in the response. The value of the modal participation factor depends on how
the mode shapes are normalized, and so will not be given here, however the values are smaller
for the higher modes with the result that the displacements for the higher modes are generally
smaller than those of the first mode. The modal displacements are presented in Table A-4 (to
three decimal places, which is why some values are shown as zero) and plotted in Figure A-9
for the first two modes as well as the RSS value. In this example, the influence of the two
highest modes is very small and has been omitted from the diagram. It is difficult to distinguish
between the first mode displacements and the RSS displacements in Figure A-9; this is
characteristic of structures with periods less than about 1 second, such as would be the case for
most masonry structures.

Figure A-9. Modal displacements.

9/1/2018 A-12
Modal analysis gives the modal shears and bending moments in each member and these
values can be used to generate the shear and moment diagrams. These are summarized in
Tables A-5 and A-6 and are graphically presented in Figure A-10. Only the results from the first
two modes are shown as the higher modes contribute very little to the response. Except for
some contribution to the shears, the second mode is insignificant in contributing to the total
values calculated using the RSS method.

a) b)
Figure A-10. Modal analysis results: a) shear forces; b) bending moments.
The inertia force at each floor for each mode can be determined by taking the difference
between the shear force above and below the floor in question. Modal inertia forces along with
the RSS values are summarized in Table A-7, and show that the higher modes at some levels
contribute more than the first mode. Note that the sum of the inertia forces for each mode is
equal to the base shear for that mode. However, the sum of the RSS values of the floor forces
at each level is 6284 kN (obtained by adding values for storeys 1 to 4 in the last column of the
table); this is not equal to the total base shear of 4468 kN found by taking the RSS of the base
shears in each mode (see Table A-3). This demonstrates the key rule in combining modal
responses: only primary quantities from each mode should be combined. For example, if
the designer is interested in the shear force diagram for the structure, it is necessary to find the
shear forces in each mode and then combine these modal quantities using the RSS method. It
is incorrect to find the total floor forces at each level from the RSS of individual modal values,
and then use these total forces to draw the shear diagram. Even interstorey drift ratios, defined
as the difference in the displacement from one floor to the next divided by the storey height,
should be calculated for each mode and then combined using the RSS procedure. It would be
incorrect to divide the total floor displacements by the storey height; although in this example
since the deflection is almost entirely given by the first mode, this approach would be very close
to that found using the RSS method.

One of the disadvantages of modal analysis is that the signs of the forces are lost in the RSS
procedure and so equilibrium of the final force system is not satisfied. Equilibrium is satisfied in
each mode, but this is lost in the procedure to combine modal quantities since each quantity is
squared. That is why it is important to determine quantities of interest by combining only the
original modal values.

9/1/2018 A-13
A.4.4. Comparison of static and modal analysis results
The equivalent static force analysis procedure, which will be presented in more detail in Section
1.6, has been applied to the four-storey structure described above for the spectrum shown in
Figure A-8. Table A-8 compares the results of the two types of analyses. It can be seen that
both the base shear and moment given by the modal analysis method is about 75% of that
given by the static method. This occurs with short period MDOF structures that respond in
essentially the first mode because the modal mass of the first mode for walls is about 70 to 80%
of the total mass. The top displacement from the modal analysis is 78% of the static
displacement, nearly the same as the ratio of the base moments; this would be expected given
that the deflection is mostly tied to the moment.

If the structure is a single-storey, SDOF system, the two analyses methods will give the same
result. But for MDOF systems, such as two-storey or higher buildings, dynamic analysis will
generally result in smaller forces and displacements than the static procedure.

The floor forces from the two analyses are quite different. The floor forces in the upper storeys
obtained by modal analysis are less than the static forces, but in the lower storeys, a reverse
trend can be observed. The reason for this is the contribution of the higher modes to the floor
forces. It can be seen in Table A-7, that at the 2nd storey, the second mode contribution is the
largest of all the modes. To ensure the required safety level when seismic design is performed
using the equivalent static analysis procedure, NBC 2015 seismic provisions (e.g. Clause
4.1.8.15) provides additional guidance on the level of floor forces to be used in connecting the
floors to the lateral load resisting elements.

Table A-1. Modal Periods and Masses

Period Modal mass/


Mode
(sec) Total mass
1 0.400 0.696
2 0.062 0.210
3 0.022 0.070
4 0.012 0.024
Sum 1.000

Table A-2. Mode Shapes


Mode Shapes
Storey
1st mode 2nd mode 3rd mode 4th mode
0 0.000 0.000 0.000 0.000
1 0.093 0.505 1.000 -1.000
2 0.328 1.000 0.334 0.969
3 0.647 0.544 -0.972 -0.619
4 1.000 -0.727 0.427 0.175
Note: mode shapes are normalized to a maximum of 1

9/1/2018 A-14
Table A-3. Spectral Accelerations, Sa, and Base Shears
Spectral Base
Period Modal mass /
Mode Acceleration Shear
(sec) Total mass
Sa (g) (kN)
1 0.400 0.74 0.696 4127
2 0.062 0.96 0.210 1617
3 0.022 0.96 0.070 534
4 0.012 0.96 0.024 184
Total base shear ABS 6462
Total base shear RSS 4468
Note: total weight = 8000 kN

Table A-4. Modal Displacements


Modal Displacements (cm)
Storey st
RSS
1 mode 2nd mode 3rd mode 4th mode
Base 0.000 0.000 0.000 0.000 0.00
1 0.367 0.021 0.002 0.000 0.37
2 1.300 0.042 0.001 0.000 1.30
3 2.564 0.023 -0.002 0.000 2.56
4 3.963 -0.031 0.001 0.000 3.96

Table A-5. Modal Shear Forces


Shear Forces (kN)
Storey st nd
RSS
1 mode 2 mode 3rd mode 4th mode
0-1 4127 1617 534 -184 4468
1-2 3942 999 -143 204 4074
2-3 3287 -224 -369 -172 3320
3-4 1996 -888 289 68 2205

Table A-6. Modal Bending Moments


Bending Moments (kNm)
Storey st
RSS
1 mode 2nd mode 3rd mode 4th mode
Base 40053 -4511 -931 255 40320
1 27675 339 670 -298 27686
2 15849 3335 240 313 16201
3 5988 2665 -867 -204 6614
4 0 0 0 0 0

9/1/2018 A-15
Table A-7. Modal Inertia Forces (Floor Forces)
Floor Forces (kN)
Storey RSS
1st mode 2nd mode 3rd mode 4th mode
1 185 618 677 -388 1012
2 655 1223 226 376 1455
3 1291 665 -658 -240 1612
4 1996 -888 289 68 2205
Sum 4127 1617 534 -184 4468

Table A-8. Comparison of Static and Dynamic Analyses Results


Storey Shear Forces Floor Forces Moments Deflections
(kN) (kN) (kNm) (cm)
Static Modal(1) Static Modal(2) Static Modal(3) Static Modal(4)
Base 0 0 53280 40320 0 0
5920 4468
1 592 1012 35520 27686 0.48 0.37
5328 4074
2 1184 1455 19536 16201 1.70 1.30
4144 3320
3 1776 1612 7104 6614 3.32 2.56
2368 2205
4 2368 2205 0 0 5.11 3.96
Notes: (1) see Table A-5, last column
(2) see Table A-7, last column;
(3) see Table A-6, last column;
(4) see Table A-4, last column.

9/1/2018 A-16
TABLE OF CONTENTS – APPENDIX B

B RELEVANT RESEARCH STUDIES AND CODE BACKGROUND B-2

B.1 Shear/Diagonal Tension Resistance B-2

B.2 Sliding Shear Resistance B-8

B.3 Ductile Seismic Response of Reinforced Masonry Shear Walls B-12

B.4 Wall Height-to-Thickness Ratio Restrictions B-16

9/1/2018 B-1
B Relevant Research Studies and Code Background
This appendix contains additional background material relevant to the aspects of masonry
design discussed in Chapter 2. Findings of some relevant research studies, as well as the
discussion on provisions of masonry design codes from other countries, are included. This
information may be useful to readers interested in gaining a more detailed insight into the
subject. However, it should be noted that designers may use alternative design provisions in
situations where CSA S304 is silent on a specific issue. The design provisions contained in
design standards from other countries cannot supersede the provisions of pertinent Canadian
standards.

B.1 Shear/Diagonal Tension Resistance

The CSA S304 shear strength design equation for RM shear walls was first included in the 1994
version of the standard (CSA S304.1-94) and it is largely based on the research performed in
1970s and 1980s, e.g. research program by the US-Japan Joint Technical Coordinating
Committee for Masonry Research (TCCMAR). Numerous experimental studies on RM shear
walls subjected to reversed cyclic loading conducted since the 1990’s provide additional data for
developing new or revised shear strength design equations.

The CSA S304 shear strength equation was evaluated by several researchers, including Seif
ElDin and Galal (2015a); El-Dakhakhni et al. (2013); Davis et al. (2010); Voon and Ingham
(2007). Davis et al. (2010) compared the estimated shear strength predictions based on 8
different code expressions (including the CSA S304.1-04) with the results from 56 tests of fully
grouted RM shear walls with shear dominated response. The average ratio of the test strength
to the estimated strength for the CSA S304 expression was 1.50 with a Coefficient of Variation
(COV) of 0.15; this is considered a rather conservative prediction.

El-Dakhakhni et al. (2013) tested 8 fully grouted cantilever RM shear wall specimens with shear
dominated behaviour subjected to reversed cyclic loading. The specimens were squat walls with
aspect ratio ranging from 0.6 to 1.5, were characterized by horizontal reinforcement ratios of
0.07 to 0.13%, and the level of applied axial stress varied from 0 to approximately 0.08xf’m. The
study examined the effectiveness of design shear strength expressions included in the
Canadian (CSA S304.1-04), US (TMS 402/ACI 530/ASCE 5-11), New Zealand (NZS
4230:2004) and European (Eurocode 6) masonry design codes. The results demonstrated that
the CSA S304.1-04 produced the most conservative predictions of all the codes (mean
experimental/calculated ratio = 1.51 and COV= 18.1%). Shear strength predictions based on
international masonry codes, especially the US TMS 402/602 code (mean= 1.14 COV= 12.7%)
and New Zealand code NZS 4230:2004 (mean= 1.13 COV= 16.9%) gave a better fit of the
experimental results.

El-Dakhakhni et al. (2013) also observed that the shear strength expression of the Canadian
concrete design standard CSA A23.3-04, based on the Simplified Modified Compression Field
Theory (SMCFT) approach (Bentz et al. 2006), gave the most accurate prediction of shear
strength for squat walls (mean= 1.06 COV= 10.8%). The underlying theory is the Modified
Compression Field Theory developed in the 1980s (Vecchio and Collins,1986), which has been
referred to as the General Method for Shear Design of RC flexural members in Canada (CSA
A23.3-04). The same approach was adopted for the design of RM beams in Canada in CSA
S304-14 (Cl.11.3.4). The design equations are similar to CSA A23.3-04, but the input parameter
values were calibrated for masonry design purposes. Also, a new parameter Kb has been

9/1/2018 B-2
introduced to take into account the level of grouting and type of masonry units. This is based on
the research by Sarhat and Sherwood (2010; 2013), which included the results of their own
experimental studies and a survey of the experimental data by other researchers.

The New Zealand Masonry Standard NZS 4230:2004 (SANZ, 2004) states that the axial load
contribution to masonry shear resistance in squat shear walls is equal to 0.9 N tan  . This
contribution results from a diagonal strut mechanism, which is based on an assumption that
axial compression load N forms a compression strut at an angle  to the vertical axis (see
Figure B-1). The axial load must be transmitted through the flexural compression zone, while
the horizontal component of the strut force resists the applied shear force (Priestley et al.,
1994). This model implies that the shear strength of squat walls under axial loads should be
greater than that of more slender walls, and higher than that prescribed in CSA S304-14.
According to this model, the axial load contribution is limited to N  0.1 f m Ag .

Figure B-1. Contribution of axial load to wall shear strength (reproduced from NZS 4230:2004
with the permission of Standards New Zealand under License 000725).

The shear strength equation in the US masonry design code TMS 402/602-16 (previous
versions were labelled as TMS 402/ACI 530/ASCE 5) was derived from research dating back to
the 1980s (Shing et al. 1990a; 1990 b). The equation has been evaluated by several
researchers, including Alogla et al. (2014); Davis et al. (2010); and Voon and Ingham (2007).
Davis et al. (2010) compared the estimated shear strength predictions based on the TMS
402/602 expression with the results from 56 tests of fully grouted RM shear walls with a shear
dominated response. The average ratio of the test strength to the estimated strength was 1.17
with a COV of 0.15, indicating that the expression is somewhat conservative. Alogla et al.
(2014) also examined the TMS 402/602 shear strength expression predictions for more than 60
walls from literature. It was observed that the shear strength calculated using the TMS 402/602
design expression overestimated the shear strength of the examined walls by about 10%.

Several design factors influence the shear/diagonal tension resistance of RM walls. A brief
overview of the available experimental research evidence on RM shear walls subjected to
reversed cyclic loading related to these factors is discussed below. El-Dakhakhni and Ashour
(2017) performed a detailed review of past experimental studies on the subject.

9/1/2018 B-3
Axial compression:

An experimental study on 16 fully grouted RM wall specimens examined the effect of axial
stress on the wall’s shear resistance (Shing et al., 1989). The axial stress ranged from 0 to
approximately 0.1xf’m. The results indicated that the load at the first diagonal crack increased
with the applied axial load. The study also demonstrated that an increasing axial load could
result in a change in the failure mechanism from a flexural/shear mode to a brittle shear mode.

An experimental study on RM wall specimens by Voon and Ingham (2006) showed that a
relatively moderate increase in axial compression stress level from 0 to 0.025xf’m resulted in an
increase in the maximum wall shear resistance of more than 20%. However, RM walls
subjected to higher axial compression had a reduced post-cracking deformation capacity,
resulting in a more brittle response. Ibrahim and Suter (1999) tested 5 squat RM shear walls
under reversed cyclic loading (aspect ratio ranged from 0.47 to 1.0) and observed that the level
of applied axial stress has a significant effect on the shear capacity.

Wall aspect ratio (squat shear walls):

The findings of several experimental studies, e.g. Matsumura (1987), Okamoto et al. (1987),
and Voon (2007), confirmed that RM walls with lower aspect ratios exhibited shear strengths
that were larger than more slender masonry walls. The researchers concluded that the shear
strength enhancement was due to the more prominent role of arching action in RM walls with
low aspect ratios, in which shear was mainly resisted by compression struts (see Figure 2-16a).
Voon and Ingham (2006) reported that the shear resistance decreased by 15% when the wall
aspect ratio increased from 1.0 to 2.0. A squat wall specimen with an aspect ratio of
approximately 0.6 showed a significant increase in shear resistance (by over 100%) compared
to an otherwise similar specimen with an aspect ratio of 1.0. The findings of an experimental
study by Okamoto et al. (1987) confirmed that the wall shear strength increased by 20 % when
the aspect ratio decreased from 2.3 to 1.6, and by 30 % when aspect ratio decreased from 2.3
to 0.9. A study on partially grouted RM walls by Schultz (1996) showed that a decrease in the
wall aspect ratio was reported to have a beneficial effect on the shear resistance, that is, squat
walls are expected to have larger shear resistance than flexural walls of the same height.
However, squat wall specimens also showed a reduced deformation capacity and increased
strength deterioration.

A few studies on RM squat shear walls subjected to reversed cyclic loading were performed in
Canada (Seif ElDin and Galal, 2015b; 2016a; 2016b; 2017; El-Dakhakhni et al., 2013). The
results confirmed the findings of other studies with regard to the shear strength of squat RM
shear walls.

Horizontal reinforcement:

Shing et al. (1989) concluded that horizontal reinforcement influences the post-cracking
response of RM walls. The study included 8 walls that failed in a shear dominated mode. and
had horizontal reinforcement ratios ranging from 0.12 to 0.22 %. The onset of cracking
(occurrence of the first major diagonal crack) depends primarily on the tensile strength of the
masonry and the applied axial load. However, increasing the amount of horizontal reinforcement
caused a change in the failure mechanism from a brittle shear mode to a ductile flexural mode.

Sveinsson et al. (1985) tested 10 RM piers (a double curvature loading condition) and varied the
amount of horizontal reinforcement from 0.075 to 0.394%. They concluded that the horizontal

9/1/2018 B-4
reinforcement was effective in increasing shear strength, but higher amounts of reinforcement
did not correspond to a proportional gain in strength. For example, a 16% increase in the shear
strength was observed in a specimen which had twice the amount of horizontal reinforcing bars
compared to an otherwise similar specimen.

Shear reinforcement in RM shear walls does not seem to be as effective as in RC shear walls. A
possible explanation is that the reinforcing bars located where the inclined crack crosses near
the end of the bar are unable to develop their full yield strength in the masonry walls. To
account for this phenomenon, the New Zealand Masonry Standard NZS 4230:2004 (SANZ,
2004) prescribes a coefficient of 0.8 in the Vs equation, while CSA S304-14 uses a 0.6 factor.
This phenomenon is particularly pronounced in short walls, where it is likely that the length of
the shear reinforcement is insufficient to fully develop its yield strength.

Seif ElDin and Galal (2015b) tested 9 squat RM walls under quasi-static cyclic loading. Contrary
to the previous experimental studies, they observed that the horizontal reinforcement
contributes to the wall shear resistance with its full yield capacity (there is no reduction
coefficient as discussed above). This can be explained by the redistribution in the shear
resistance between the reinforcement and the masonry, especially at high ductility demands.
Most previous researchers quantified shear contribution of reinforcement based on the
difference between the shear capacities of specimens with different transverse reinforcement
ratios.

It appears that horizontal reinforcement in RM shear walls does not have as good anchorage as
the corresponding reinforcement in RC shear walls. Anderson and Priestley (1992) have noted
that straight bars or 90 hooks were used in some experimental studies (see Figure B-2a),
whereas the horizontal reinforcement in RC shear walls is usually anchored in a more effective
way, such as by 180 hooks. The type and extent of anchorage are expected to influence the
effectiveness of shear reinforcement. Sveinsson et al. (1985) tested 10 fully grouted RM piers
and studied (among other factors) the effect of anchorage conditions in horizontal reinforcement
(90 versus 180 hooks). They recommended the use of 180 hooked end anchorage for
horizontal reinforcement because it produced better energy dissipation, and enabled the bars to
develop their full tensile strength This is particularly true for shorter walls/piers.

Seif ElDin and Galal (2016a) tested 3 squat RM wall specimens with shear dominant behaviour
under reversed cyclic loading. The specimens were identical, except for the end anchorage of
the horizontal reinforcing bars: the first specimen had 180 hooks, the second one 90 hooks,
and the third one had straight bars (no hooks). The results showed that the specimen with 180
hooks provided the most effective anchorage and attained the largest shear capacity and
displacement ductility, while the specimen with straight bars attained the smallest shear
capacity and displacement ductility. However, the difference in the strength values was not
significant (it was within 10%). The most significant difference was in the post-peak behaviour.
The specimen with straight bars showed the most rapid post-peak degradation of the lateral
load resistance. The 180 hooks proved to be effective in providing confinement for the vertical
end bars in the wall, while the 90 hooks were less effective. For that reason, displacement
ductility of the specimen with 180 hooks (4.2) was higher than the specimen with 90 hooks
(3.9) and the one with straight bars (3.6). This difference again indicates the superior ductility
potential of the 180 end hooks, but the other anchorage conditions may be acceptable in some
cases. The researchers recommended the use of horizontal reinforcing bars with 90 hooks for
masonry structures located in regions of low to moderate seismic hazard, and/or outside the
plastic hinge regions in ductile shear walls.

9/1/2018 B-5
Vertical reinforcement:

Anderson and Priestley (1992) found that shear strength didn’t show any correlation with the
vertical reinforcement ratio, hence the CSA S304 shear design equation ignore the effect of
vertical reinforcement. However, according to some researchers (Shing et al., 1990; Tomazevic,
1999; Voon, 2007), a fraction of the wall shear resistance can be attributed to the presence of
vertical reinforcement. Dowel action in vertical reinforcing bars enables shear transfer across a
diagonal crack by the localized kinking in reinforcing bars due to their relative displacement (see
Figure B-2b) (note that compression kinks cancel out some of the tension kinks). However, once
the vertical reinforcement yields, as it would in the plastic hinge zone of ductile walls, its
contribution to the shear resistance drops significantly and could be ignored.

Figure B-2. Wall reinforcement contributing to shear resistance: a) horizontal reinforcement


acting in tension; b) dowel action in vertical reinforcement (Tomazevic, 1999, reproduced by
permission of the Imperial College Press).

Ductility:

Experimental studies on RM shear walls with shear dominant behaviour (aspect ratio less than
2.0) have demonstrated that significant levels of ductility and energy dissipation capacity are
possible in these walls (Sveinsson et al. 1985; Shing et al. 1989; Voon and Ingham 2006; El-
Dakhakhni et al. 2013). Shing et al. (1989) observed that the displacement ductility ratio tends
to increase with an increase of axial load for the shear dominated specimens. They attributed
the increased ductility level to the aggregate interlock forces which are enhanced by the
increase of axial load.

It has been recognized that shear degradation at higher ductility demands occurs in shear-
dominated RM walls. In their empirical equation which estimates the shear strength of RM shear
walls, Anderson and Priestley (1992) proposed factor k to account for the degradation of the
shear resistance provided by masonry for the inelastic response when the displacement ductility
ratio increases from 2.0 to 4.0. The value decreases linearly from 1.0 to 0 as the displacement
ductility ratio increases from 2.0 to 4.0.

9/1/2018 B-6
Grouting:

Experimental studies have reported a significant reduction in the shear resistance of partially
grouted walls compared to otherwise identical fully grouted walls. Brzev (2011) performed a
review of available experimental data related to the subject. The review included 29 partially
grouted RM wall specimens tested in the period from 1978 to 2010, including Nolph (2010);
Nolph and ElGawady (2012); Elmapruk (2010); Minae et al. (2010); Maleki (2008); Maleki et al.
(2009); Voon (2007a); Schultz (1996); and Chen et al. (1978). Most specimens (24 out of 29)
were squat RM walls and had a horizontal reinforcement ratio of 0.07% or higher and 180
hooks. All specimens had a vertical reinforcement ratio of 0.07% or higher, while 15 out of 29
specimens had a vertical reinforcement ratio of 0.3% or higher.

Lateral load resisting mechanisms for lightly reinforced partially grouted RM shear walls are
significantly different than for fully grouted walls. Research evidence related to the seismic
response of partially grouted walls consists primarily of experimental studies where individual
wall specimens were subjected to quasi-static cyclic loading, although there are also a few
shake-table studies.

Most research studies on specimens subjected to quasi-static cyclic loading report shear
dominated mechanism of seismic response characterized by stair-stepped and/or diagonal
tension cracks in the masonry panels enclosed by grouted bond beams and vertical cells. These
cracks are indicative of the formation of compression struts within the panel. The failure is often
accompanied by spalling of face shells in the block units (Nolph, 2010).

In general, the response of tested specimens to the cyclic loads was reasonably stable. None of
the specimens displayed a sudden failure, and the resistance gradually deteriorated with
progressively increasing cyclic loading.

Most specimens achieved a displacement ductility ratio of 2.0 or higher, except for the
specimens tested by Nolph (2010) and Elmapruk (2010), which were characterized by relatively
high vertical reinforcement ratios (0.46% for the Nolph specimens and 0.33% for the Elmapruk
specimens). It was observed that the displacement ductility ratio decreased with an increase in
the vertical reinforcement ratio. The specimens tested by Voon (2007a) also showed a ductility
ratio of less than 2.0, but these specimens had no horizontal reinforcement.

Schultz (1996) tested a series of 6 partially grouted RM wall specimens under in-plane cyclic
loads. Only the outermost vertical cores and a single course bond beam at midheight were
grouted. The mechanism of shear resistance in the tested walls was characterized by the
development of vertical cracks between the ungrouted and grouted masonry due to stress
concentrations or planes of weakness (this mechanism is different from the one expected to
develop in solidly grouted RM walls). It was also reported that an increase in horizontal
reinforcement ratio did not have a significant effect on the overall shear resistance.

An experimental study by Voon and Ingham (2006) showed that the shear strength of a solidly
grouted wall specimen was approximately 110% higher than an otherwise identical specimen
with 30% grouted cores. Also, the specimen with 55% grouted cores had a shear strength more
than 50% higher than the specimen with 30% grouted cores. However, the difference
decreases when the shear stress is compared using the net wall area.

9/1/2018 B-7
Ingham et al. (2001) reported the results of an experimental study on 12 full-scale RM squat
wall specimens subjected to in-plane cyclic lateral loading (aspect ratios ranged from 0.57 to
1.33). Of the twelve specimens, nine were partially grouted, and three were fully grouted. The
walls were designed to fail in the diagonal tension shear mode. The test results showed that the
fully grouted RM wall specimens demonstrated significantly higher displacement ductility (on the
order of 6.0) than the displacement ductility of otherwise identical partially grouted specimens
(about 4.0) It should be noted that all partially grouted specimens achieved a displacement
ductility of 2.0 or higher. A possible reason for the higher ductility in the fully grouted RM wall
specimens is that they ultimately failed in a sliding shear mode, which is characterized by large
deformations at the base of the wall. The partially grouted specimens failed in the diagonal
tension mode. Force-displacement responses for a partially grouted Wall 2 and a fully grouted
Wall 3 specimen are shown in Figure B-3 (the specimens were otherwise similar, except for the
grouting pattern).

Figure B-3. Force-displacement responses for partially grouted (left) and fully grouted (right)
wall specimens (Ingham et al., 2001, reproduced by permission of the Masonry Society).

B.2 Sliding Shear Resistance

Sliding shear resistance according to the CSA S304-14 standard has been determined based
on friction resistance from Coulomb’s Law, as discussed in Section 2.3.3. However, a sliding
shear mechanism is also characterized by sliding displacements along the sliding interface
(usually base of the wall). In long walls with openings consisting of several interconnected piers,
sliding movements at the base of one pier might cause damage in the adjacent piers. However,
current international masonry design codes, including CSA S304-14, do not contain provisions
for estimating sliding displacements in the walls or corresponding displacement limits. Centeno
(2015) studied sliding failure mechanisms in RM shear walls and estimated sliding
displacements due to lateral loading. He proposed a Sliding Shear Behavior (SSB) method for
estimating the base sliding displacements in RM shear walls (Centeno, 2015; Centeno et al.,
2015). This section summarizes the method, which can be applied through a step-by-step
process. The objective of the process is to determine: 1) the wall’s yield mechanism, and 2) the
magnitude of sliding displacements that occur in that mechanism. There are two principal yield
mechanisms associated with sliding shear (Figure B-4): a) a sliding shear mechanism and b) a
combined flexural-sliding shear mechanism. The sliding shear mechanism occurs when the
lateral force, V, is equal to or greater than the sliding shear resistance of the RM wall, where the
sliding displacements develop at the base of the wall. The combined flexural-sliding shear
mechanism occurs when the RM wall yields in flexure and forms an open flexural crack along

9/1/2018 B-8
the wall length. Inelastic displacements in the wall are equal to the sum of flexural and shear
displacements.

Figure B-4. Yield mechanisms in RM shear walls subjected to monotonic lateral loading: a)
sliding shear mechanism and b) flexural yield mechanism (Centeno, 2015).

For displacement estimation purposes, Centeno (2015) identified three yield mechanisms that
lead to sliding displacements: i) Sliding Shear (SS) mechanism, ii) Combined Flexural-Sliding
Shear (CFSS) mechanism, and iii) Sliding Failure (SF) mechanism. These mechanisms are
based on the two mechanisms illustrated in Figure B-4. The SS mechanism is illustrated in
Figure B-4a), while the remaining two mechanisms (CFSS and SF) are variants of mechanism
shown in Figure B-4b). In RM walls that experience a SS mechanism, sliding displacements
occur when an applied lateral force exceeds the wall’s sliding shear resistance. In the walls that
experience a CFSS or a SF mechanism, sliding displacements are the result of dowel
deformations that occur in order for dowel action to transfer shear across an open flexural crack
during cyclic loading. In a CFSS mechanism, displacements are elastic but influenced by
degradation in dowel action shear stiffness, while in a SF mechanism, the displacements are
inelastic and occur when the applied shear force exceeds the dowel action yield resistance.

The procedure for estimating sliding displacements according to the SSB method is presented
below.

Part 1: Determine the Wall’s Yield Mechanism

Step 1: Determine the plastic moment resistance, Mp, and its corresponding lateral force
resistance, VFl.

Step 2: Establish the Upper Bound Sliding Shear Resistance, :


(B.1)
, where (B.2)

(B.3)

9/1/2018 B-9
(B.4)

(B.5)

where:
: masonry cover s: rebar spacing
: masonry grout compression strength fy: reinforcing steel yield stress
(MPa)
As: total area of reinforcing steel : axial compression force
µFr: friction coefficient, (µFr = 0.6) c: depth of compression zone
H: wall height L: wall length
: height to length aspect ratio
: friction force due axial compression : friction force due to flexural
compression
(upper bound)
: dowel action yield resistance CDA: dowel action strength coefficient

Step 3: Determine if the yield mechanism is a Sliding Shear (SS) Mechanism:


If < , then yield mechanism is Sliding Shear Mechanism. Continue to Part II, Step A1.
If ≥ , then yield mechanism is not Sliding Shear Mechanism. Continue to Step 4.

Step 4: Calculate the overturning moment, Mo, and corresponding lateral force, Vo, required to
close flexural crack during cyclic loading:
4.1: Determine the overturning moment, Mo:
(B.6)

(B.7)
where:
Mo: overturning moment to close flexural crack
CM: overturning moment coefficient
Vo: lateral force to close flexural crack

Step 5: Determine if yield mechanism is Sliding Failure Mechanism


If < , then yield mechanism is Sliding Failure Mechanism. Must increase the wall’s dowel
resistance, , and return to step 1.
If > , then yield mechanism is not Sliding Failure Mechanism. Continue to Step 6.

Step 6: Determine if yield mechanism is a Combined Flexural Sliding Shear (CFSS)


Mechanism.
6.1: Calculate the upper limit aspect ratio, TAR2, for which a wall develops a CFSS mechanism.
(B.8)
If H/L < TAR2 then yield mechanism is CFSS Mechanism. Continue to Part II, Step B1.

9/1/2018 B-10
If H/L ≥ TAR2 then yield mechanism is a Flexural Mechanism. Sliding displacements in the wall
design will be small. If necessary, the sliding displacements can be measured by continuing to
Part II, Step B1.

Part II: Estimate the Sliding Displacements

Step A: Estimate sliding displacements for a SS mechanism

A1: Calculate the upper limit aspect ratio, TAR1, for which a wall develops a SS mechanism.
TAR1 = H/L (when ) (B.9)
(Note: Calculating TAR1 requires trying multiple values of H/L until finding the aspect ratio that
meets the condition in equation B.9)
A2: Calculate the friction from flexural compression, ,
This is a correction of the friction force component that corresponds to flexural yielding, because
in a wall that develops a sliding shear mechanism not all of the tension reinforcement will reach
its yielding stress due to flexure. Therefore, the friction force, FrFl, is only a fraction of the upper
bound friction force, , determined in step 2.
(B.10)
A3: Determine sliding shear resistance, VSS, due to a SS mechanism:
(B.11)

A4: Calculate wall lateral stiffness, ,


Following the recommended empirical equation by Shing et al. (1990) for the lateral stiffness of
a wall with a shear-dominant response:
(B.12)

(B.13)

where:
: elastic shear stiffness : Elastic Modulus of Masonry
: post-cracking shear stiffness : Poisson ratio, (for Masonry, = 0.2)
t: wall thickness

A5: Sliding Displacement Equation for SS Mechanism,


, when (B.14)
where:
: wall base sliding displacement
µ: displacement ductility ratio

Step B: Estimate sliding displacements for a CFSS mechanism

B1: Determine Triggering aspect ratios: TAR1, TAR2 and TAR3.


TAR1 = H/L when (B.9)

(B.8)

TAR3 = H/L when (B.15)

9/1/2018 B-11
Note that calculation of TAR1 and TAR3 requires trying multiple values of H/L until finding the
aspect ratio that meets the condition in equations B.9 and B.15, respectively)

B2: Calculate dowel action secant stiffness coefficient, Ck.


(B.16a)

(B.17b)
where:
µ: displacement ductility ratio
B3: Determine dowel action yield stiffness, k DA.

(B.18)

Note: f’g (MPa), db (mm) (B.19)

B4: Calculate base sliding displacement, ∆Base.


(B.20)

B.3 Ductile Seismic Response of Reinforced Masonry Shear Walls


A prime consideration in seismic design is the need to have a structure that is capable of
deforming in a ductile manner when subjected to several cycles of lateral loading well into the
inelastic range. This section explains a few key terms related to ductile seismic response,
including ductility ratio, curvature, plastic hinge, etc. It is important for a structural designer to
have a good understanding of these concepts before proceeding with the seismic design and
detailing of ductile masonry walls according to CSA S304-14. In particular, the content of this
section is related to the ductility check for RM shear walls discussed in Section 2.6.3.

Ductility is a measure of the capacity of a structure or a member to undergo deformation beyond


yield level, while maintaining most of its load-carrying capacity. Ductile structural members are
able to absorb and dissipate earthquake energy by inelastic (plastic) deformations that are
usually associated with permanent structural damage. These inelastic deformations are
concentrated mainly in regions called plastic hinges. In general, plastic hinges develop in shear
walls responding in the flexural mode and are typically formed at their base. An example of a
plastic hinge formed in a RM wall subjected to seismic loading is shown in Figure 2-8a. The
concept of ductility and ductile seismic response was introduced in Section 1.4.3.

A common way to quantify ductility in a structure is through the displacement ductility ratio   .
This is the ratio of the maximum lateral displacement experienced by the structure at the
ultimate (  u ), to the displacement at the onset of inelastic response (  y ) (see Figure 1-5c).

u
 
y

9/1/2018 B-12
Next, the concept of curvature will be explained by an example of a RM shear wall subjected to
bending due to a shear force applied at the top, as shown in Figure B-5a. Consider a wall
segment ABCD of unit height. This segment deforms due to bending moments, so sections AB
and CD rotate by a certain angle relative to their original horizontal position (these deformed
sections are denoted as A’B’ and C’D’). Rotation between the ends of the segment defines the
curvature  , as shown in Figure B-5b. Curvature represents relative section rotations per unit
length. It should be noted that curvature is directly proportional to the bending moment at the
wall section under consideration, if the section remains elastic.

Consider any section CD that undergoes curvature  , as shown in Figure B-5c. Strain
distribution along the wall section is defined by the product of curvature and the distance from
the neutral axis, located by the depth c . The maximum compressive strain in masonry  m is
given by

m   c

Figure B-5. Curvature in a shear wall subjected to flexure: a) wall elevation; b) deformed wall
segment ABCD; c) strain distribution along the section CD.
For the seismic design of RM walls, it is of interest to determine curvatures at the following two
stages: the onset of steel yielding and at the ultimate stage. Consider a RM wall section
subjected to axial load and bending shown in Figure B-6a.
Yield curvature  y corresponds to the onset of yielding characterized by tensile yield strain  y
developed in the end rebars, as shown in Figure B-6b, where

y
y 
lw  d   c
Ultimate curvature  u corresponds to the ultimate stage, when the maximum masonry
compressive strain  m has been reached. The maximum  m value has been limited to 0.0025
by CSA S304-14 (see Figure B-6c) to prevent damage to the outer blocks in the plastic hinge

9/1/2018 B-13
region. Note that the neutral axis depth c is going to decrease as more of the reinforcement has
yielded (see Figure B-6c).

Figure B-6. Curvature in a RM wall section: a) wall cross section; b) yield curvature; c) ultimate
curvature; d) moment-curvature relationship.
The curvature value depends on the load level, the section geometry, the amount and
distribution of reinforcement, and the mechanical properties of steel and masonry. An actual
moment-curvature relationship for ductile sections is nonlinear, however it is usually idealized by
elastic-plastic (bilinear) relationship, as shown in Figure B-6d.

Once the curvatures at the critical stages have been determined, the curvature ductility ratio 
can be found as follows
u
 
y

When the curvature distribution along a structural member (e.g. shear wall) is defined, rotations
and deflections can be calculated by integrating the curvatures along the member. This can be
accomplished in several ways, including the moment area method.

9/1/2018 B-14
Rotations and deflections in a masonry shear wall at the ultimate state can be determined
following the approach outlined above. Consider a cantilevered shear wall of length l w and
height hw , and the plastic hinge length l p (see Figure B-7a). The wall is subjected to a seismic
shear force at the top, which results in a corresponding bending moment diagram as shown in
Figure B-7b. The curvature diagram shown in Figure B-7c has two distinct portions: an elastic
portion, with the maximum curvature equal to the yield curvature  y , and the plastic portion with
the maximum curvature equal to the ultimate curvature  u . Note that the elastic portion of the
curvature diagram has the same shape as the bending moment diagram (since the curvatures
and bending moments are directly proportional). The actual curvature distribution in the plastic
region varies in a nonlinear manner, as shown in Figure B-7c. For design purposes, the
curvature can be taken as constant over the plastic hinge length l p (note that the areas under
the actual and the equivalent plastic curvature are set to be equal). The elastic rotation  e and
the plastic rotation  p are presented in Figure B-7d. The plastic rotation can be determined as
the area of the equivalent rectangle of width  u   y and height l p , as shown in Figure B-7c.
These rotations can be calculated from the curvature diagram as follows:
u  e   p
where
 y  hw
e 
2
 p   u   y  l p

The maximum deflection  u at the top of the wall is shown in Figure B-7d. This deflection has
two components: elastic deflection  y corresponding to the yield curvature  y , and the plastic
deflection  p due to a rigid body rotation, since bending moments do not increase once the
yielding has taken place. Deflection values can be found by taking the moment of the curvature
area around point A, as follows:

 y hw 2hw  y hw
2

y   
2 3 3

 p   u   y   l p hw  0.5l p 

u   y   p

The above equations can be used to determine the displacement ductility ratio   , in terms of
the curvature ductility  and other parameters, as follows:

u  l p  lp 
   1  3  1 1  0.5 
y  hw  hw 
Alternatively, the curvature ductility ratio  can be expressed in terms of the displacement
ductility ratio, as follows:

u hw    1
2

   1
 y 3l p hw  0.5l p 

It should be noted that   and  values are different for the same member. Once the yielding
has taken place, the deformations concentrate at the plastic hinges, so the curvature ductility 

9/1/2018 B-15
is expected to be larger than the displacement ductility   . This difference is more pronounced
in walls with larger displacement ductility ratios.

Figure B-7. Shear wall at the ultimate: a) wall elevation; b) bending moment diagram;
c) curvature diagram; d) deflections.

B.4 Wall Height-to-Thickness Ratio Restrictions


The out-of-plane wall instability of RM and RC shear walls due to in-plane lateral reversed cyclic
loading is a complex phenomenon, which has proven to be difficult to account for by means of a
rational mechanics-based approach. The out-of-plane instability of RC shear walls in multi-
storey buildings was observed in the 2010 Maule, Chile earthquake (M 8.8) (Westenenk et al.
2012) and the 2011 Christchurch, New Zealand earthquake (M 6.3) (Elwood 2013). However,
there is no evidence of out-of-plane instability for RM shear walls in past earthquakes, and
experimental research evidence is extremely limited. Azimikor et al. (2011) and Herrick (2014)
performed a literature review of past experimental studies related to this subject.

A pioneering research study on this subject was undertaken by Paulay and Priestley (1992,
1993). They concluded that a RC or RM shear wall can experience lateral instability when the
longitudinal reinforcement in its end zones is subjected to compression loads subsequent to
cycles of tensile plastic strain. Horizontal cracks form along the height of the plastic hinge region
in the wall end zone during tension load cycles, and may not fully close during subsequent
compression load cycles. Due to the presence of open cracks and the residual plastic strains in
the vertical reinforcement within the wall end zone, that zone becomes very flexible and
susceptible to significant out-of-plane displacements at low compression stress levels. It is
possible to determine the critical out-of-plane displacement beyond which instability will occur
for a specific design case. This displacement is equal to the minimum distance between the
centroid of steel and face of masonry block. For example, the critical displacement is equal to
b/2 for a wall with thickness b and one layer of longitudinal reinforcement (where a reinforcing
bar is placed in the centre of a hollow core).

9/1/2018 B-16
Paulay and Priestley (1993) developed an analytical model which offers a means to find the
minimum wall thickness required to avoid out-of-plane instability. The minimum thickness value
depends on several parameters, including the vertical reinforcement ratio, the desired curvature
and displacement ductility ratios, the plastic hinge length, and the mechanical properties of the
steel and masonry. Paulay and Priestley also performed an experimental study to confirm their
analytical model. They tested a few reinforced concrete shear wall specimens and a concrete
masonry wall specimen. The masonry wall specimen failed by out-of-plane buckling at a very
large displacement ductility   of around 14.

The application of this procedure will be illustrated on an example of a RM wall. The equation
for the critical wall thickness bc is as follows (Paulay and Priestley, 1992)

bc  0.022lw 
Curvature ductility,  , is related to displacement ductility,  , as shown in Section B.3. The
plastic hinge length l p is taken equal to hw 6 , and so the equation can be simplified as follows

  2.2   1

The displacement ductility ratio  can be considered equal to Rd prescribed by NBC 2015 for
different SFRSs (note that  values in the range from 2.0 to 3.0 are considered in this
example). By following the above procedure, it is possible to obtain the bc l w ratios
corresponding to different  values. The results are summarized in Table B-1.

For example, if the wall length l w is equal to 5,000 mm, the corresponding critical thickness bc is
equal to 150 mm for  = 2.0, or 230 mm for  = 3.0. Paulay and Priestley suggest that the
critical wall thickness should be expressed as a fraction of the wall length rather than its height.

Table B-1. Critical Wall Thickness bc Versus the Displacement Ductility Ratio 

  l w bc
2.0 2.2 31
2.5 3.3 25
3.0 4.4 22

A recent Canadian experimental program (Azimikor 2012; Robazza 2013; Azimikor et al. 2012;
2017; Robazza et al. 2017a; 2017b; 2018) demonstrated that the out-of-plane wall instability is
difficult to induce in RM shear walls at the ductility demand levels relevant for Canadian
masonry design practice. Phase 1 of the program focused on simulating the behaviour of the
wall end zones using uniaxial specimens. The purpose of the study was to understand the out-
of-plane instability phenomenon and identify key factors influencing its development. Phase 2
consisted of testing several full-scale RMSW specimens under in-plane reversed cyclic loading.
Masonry for the test specimens was laid in 50% running bond using Type S mortar for faceshell
bedding and standard Canadian concrete hollow block units.

Phase 1 consisted of testing 5 prismatic specimens with a rectangular cross-section (600 mm


length and 140 mm thickness), which were designed to simulate the end zone of a RM shear
wall (Azimikor 2012; Azimikor et al. 2012; 2017). All specimens had the same height (3.8 m),
resulting in a h/t ratio of 27. The vertical reinforcement ratio varied from 0.24% (the minimum
permissible by CSA S304.1-04) to 1.07% (the maximum practical in the masonry industry). The

9/1/2018 B-17
loading protocol consisted of reversed-cyclic uniaxial tension and compression displacement
cycles of incrementally increasing magnitude until failure. Four specimens experienced out-of-
plane instability, while the fifth specimen was a reference specimen which was subjected to
monotonic compression and experienced a compression/crushing failure. These tests had some
limitations: the specimens were isolated and were not able to simulate actual boundary
conditions along the wall height and the effect of strain gradient along the wall length. It was
concluded that the level of applied tensile strain in a wall end-zone was one of the critical factors
governing its out-of-plane stability. The maximum tensile strain that may be imposed on a
ductile RM shear wall’s end-zone could be determined, at least in part, by a kinematic
relationship between the axial strain and the out-of-plane displacement. A preliminary
mechanical model was proposed which provided a theoretical prediction of the maximum tensile
strain before an instability would take place.

Phase 2 comprised of an experimental study of 8 full-size RMSW specimens of varying h/t and
aspect (h/L) ratios, vertical and horizontal reinforcement amounts and detailing, applied axial
pre-compression, and cross-section shape (6 specimens had regular rectangular cross-
sections, while the other 2 specimens had T-shaped cross-sections) (Robazza 2013; Robazza
et al. 2017a; 2017b; 2018). The specimens were subjected to either cyclic or reversed-cyclic
loading until failure. All specimens were designed to exhibit flexure-controlled behavior
characterized by the development of high tensile strains over a distinct region of plastic hinging,
which is a theoretical prerequisite for the occurrence of out-of-plane instability. The specimens
had aspect ratios varying from 1.5 to 3.0, which were required to maintain a relatively large
plastic hinge height while still avoiding a shear failure. The specimens were designed with
relatively high h/t ratios, ranging from 21.1 to 28.6, which exceeded the maximum CSA S304.1-
04 limits for ductile RM shear walls. However, only one specimen experienced out-of-plane
displacements large enough to precipitate instability, which occurred only after the wall had
reached its ultimate shear capacity and experienced substantial degradation.

It was found that several factors may influence the out-of-plane response of RM shear
wallsubjected to in-plane loading, including ductility and tensile strain demands, applied pre-
compression levels and construction practices, as well as the effects of alternative failure
mechanisms. This research also demonstrated that the strain gradient in a RM wall is a very
important factor. This was not included in previous numerical models for out-of-plane stability in
RM or RC shear walls, which were developed exclusively based on data from testing uniaxial
specimens (e.g. Paulay and Priestley, 1993; Chai and Elayer, 1999). The estimates based on
these models may lead to overly conservative h/t requirements.

Findings of the research by Paulay and Priestley (1992; 1993) were incorporated in the seismic
design provisions for RM shear walls in New Zealand. The New Zealand masonry design
standard NZS 4230:2004 prescribes the following minimum thicknesses for limited ductility walls
(   of 2.0) and ductile walls (   of 4.0):
1. For walls up to 3 storeys high (Cl.7.4.4.1 and 7.3.3), minimum thickness t should not be
less than Ln 20 (or 0.05 Ln ), where Ln denotes clear vertical distance between lines of
effective horizontal support or clear horizontal distance between lines of effective vertical
support. Commentary to Cl.7.3.3 states that “for a given wall thickness, t , and the case
when lines of horizontal support have a clear vertical spacing of Ln  20t , then vertical lines
of support having a clear horizontal spacing of Ln  20t shall be provided.”
2. For walls more than 3 storeys high (Cl.7.4.4.1) minimum thickness t shall not be less than
Ln 13.3 (or 0.075 Ln ). However, a smaller wall thickness can be used provided that one of
the following conditions is satisfied (maximum strain in masonry  u is equal to 0.003
according to NZS 4230:2004) (see Figure 2-28):
a) c  4t or

9/1/2018 B-18
b) c  0.3l w or
c) c  6t from the inside of a wall return of a flanged wall, which has a minimum length
0.2 Ln .
The relaxed thickness requirement applies to the cases where the neutral axis depth is small,
and so the compressed area may be so small that the adjacent vertical strips of the wall will be
able to stabilize it. This is likely the case with rectangular walls subjected to low axial
compression.

Commentary to NZS 4230 Cl.7.4.4.1 states that it is considered unlikely that failure due to
lateral instability of the wall will occur in structures less than 3 storeys high, because of the rapid
reduction in flexural compression with height. This is also in line with the statement made by
Paulay (1986), that out-of-plane stability is likely to take place in walls with large plastic hinge
length (one storey or more).

Paulay and Priestley (1992) stated that “where the wall height is less than three storeys, a
greater slenderness should be acceptable. In such cases, or where inelastic flexural
deformations cannot develop, the wall thickness t need not be less than 0.05Ln ” (where
Ln denotes clear wall length between the supports).

FEMA 306 (1999) also discusses the issue of wall instability. This document also refers to the
procedure by Paulay and Priestley (1993) and provides the following recommendation for
minimum wall thickness in ductile walls (   of 4.0):

t  l w 24 or t  h 18

Note that the above requirement, which applies to the walls with displacement ductility ratio
(   ) equal to 4.0.

FEMA 306 (1999) also points out that “the lack of evidence for this type of failure in existing
structures may be due to the large number of cycles at high ductility that must be achieved –
most conventionally designed masonry walls are likely to experience other behaviour modes
such as diagonal shear before instability becomes a problem.”

9/1/2018 B-19
TABLE OF CONTENTS – APPENDIX C

C RELEVANT DESIGN BACKGROUND ............................................................................ C-2

C.1 Design for Combined Axial Load and Flexure.......................................................... C-2


C.1.1 Reinforced Masonry Walls Under In-Plane Seismic Loading ............................................... C-2
C.1.2 Reinforced Masonry Walls Under Out-of-Plane Seismic Loading ........................................ C-9

C.2 Wall Intersections and Flanged Shear Walls .......................................................... C-15


C.2.1 Effective Flange Width ........................................................................................................ C-15
C.2.2 Types of Intersections ......................................................................................................... C-16
C.2.3 Shear Resistance at the Intersections ................................................................................ C-18

C.3 Wall Stiffness Calculations ...................................................................................... C-22


C.3.1 Lateral Load Distribution ..................................................................................................... C-22
C.3.2 Wall Stiffness: Cantilever and Fixed-End Model ................................................................. C-23
C.3.3 Approximate Method for Force Distribution in Masonry Shear Walls ................................. C-24
C.3.4 Advanced Design Approaches for Reinforced Masonry Shear Walls with Openings......... C-27
C.3.5 The Effect of Cracking on Wall Stiffness ............................................................................. C-32

9/1/2018 C-1
C Relevant Design Background
This appendix contains additional information relevant for masonry design as discussed in
Chapter 2, but it is not directly related to the seismic design provisions of CSA S304-14.
Applications of the design methods and procedures presented in this appendix can be found in
Chapter 3, which contains several design examples. This appendix addresses in detail several
topics of interest to masonry designers, e.g., the calculation of in-plane wall stiffness, including
the effect of cracking, and force distribution in perforated shear walls. However, modeling and
analysis of multi-storey perforated shear walls are not covered in this document.

C.1 Design for Combined Axial Load and Flexure


C.1.1 Reinforced Masonry Walls Under In-Plane Seismic Loading

10.2

Seismic shear forces acting at floor and roof levels cause overturning bending moments in
shear walls, which reach a maximum at the base level. In general, shear walls are subjected to
the combined effects of flexure and axial gravity loads. The theory behind the design of masonry
wall sections subjected to effects of flexure and axial load is well established, and is essentially
the same as that of reinforced concrete walls. A typical reinforced masonry wall section is
shown in Figure C-1a), along with the distribution of internal forces and strains arising from the
axial load and moment. According to CSA S304-14, the strain distribution along the wall length
is based on the assumptions that the wall section remains plane and that the maximum
compressive masonry strain  m is equal to 0.003 (see Figure C-1b)). Figure C-1c) shows the
distribution of internal forces on the base of the wall, as well as the axial load, Pf and the
bending moment, M f . In the compression zone, the equivalent rectangular stress block has a
depth a , and a maximum stress intensity of 0.85  m f ' m . Note that the  factor assumes a
value of 1.0 for members subjected to compression perpendicular to the bed joints, such as
structural walls (S304-14 Cl.10.2.6). Each reinforcing bar develops an internal force (either
tension or compression) equal to the product of the factored stress and the corresponding bar
area. The internal vertical forces must be in equilibrium with Pf , and the factored moment
capacity M r can be determined by taking the sum of the moments of the internal forces around
the centroid of the section.

The following three design scenarios and the related simplified design procedures will be
discussed in this section:
1. Wall reinforcement (both concentrated and distributed) and axial load are given – find
moment capacity
2. Wall is reinforced with distributed reinforcement only – find moment capacity
3. Wall reinforcement needs to be estimated (factored bending moment and axial force are
given)

The first two are applicable for the common situations where a designer assumes the minimum
seismic reinforcement amount and desires to find its moment capacity.

Approximate design approaches that can be used to assist designers in each of these scenarios
are presented below. For detailed analysis and design procedures, the reader is referred to
Drysdale and Hamid (2005) and Hatzinikolas, Korany and Brzev (2015).

9/1/2018 C-2
Figure C-1. A reinforced masonry shear wall under the combined effects of axial load and
flexure: a) plan view cross section; b) strain distribution; c) internal force distribution.

C.1.1.1 Moment capacity for a wall section with concentrated and distributed
reinforcement

Rectangular section
A simplified wall design model is shown in Figure C-2. The wall reinforcement can be divided
into:
 Concentrated reinforcement at the ends (area Ac at each end), and
 Distributed reinforcement along the wall length (total area Ad ).
It is assumed that the concentrated wall reinforcement yields either in tension or in compression
at the wall ends. Also, it is assumed that the distributed reinforcement yields in tension.

A procedure to find the factored moment capacity M r for a shear wall with a given vertical
reinforcement (size and spacing) is outlined below.

From the equilibrium of vertical forces (see Figure C-2b)), it follows that
Pf  T1  T2  C3  C m  0 ( 1)

where
T1  C3   s f y Ac
T2   s f y Ad
Cm  0.85m f 'm t  a 
The compression zone depth, a , can be determined from equation 1 as follows
Pf   s f y Ad
a ( 2)
0.85 m f ' m t

9/1/2018 C-3
1  0.8 when f 'm  20 MPa (note that 1 value decreases when f 'm  20 MPa, as prescribed
in S304-14 Cl.10.2.6)

The neutral axis depth, c , measured from the extreme compression fibre to the point of zero
strain is given by
c  a 1
Next, the factored moment capacity, M r , can be determined by summing up the moments
around the centroid of the wall section (point O) as follows

r m w  s y c w

M  C (l  a ) 2  2  f A l 2  d ' 

 ( 3)

where d  is the distance from the extreme compression fibre to the centroid of the concentrated
compression reinforcement.

Figure C-2. A simplified design model for rectangular wall section: a) plan view cross-section
showing reinforcement; b) internal force distribution.

10.2.8
For squat shear walls, CSA S304-14 prescribes the use of a reduced effective depth d for
flexural design, i.e.
d  0.67l w  0.7 h
As a result, the moment capacity should be reduced by taking a smaller lever arm for the tensile
steel, as follows:

r m w  s y c w
 
M  C (l  a) 2   f A l 2  d '    f A d  l 2 
  s y c w 
  ( 4)

Note that the reinforcement area Ac in squat walls should be increased to provide more than
one reinforcing bar, since the end zone constitutes a larger portion of the overall wall length in
these cases.

9/1/2018 C-4
The CSA S304-14 provision for the reduced effective depth in squat walls contained in Cl.10.2.8
is intended to account for the effect of the deep beam behaviour of squat walls. This provision
makes more sense for non-seismic design, and it should not be used if the tension steel yields
in seismic conditions.

Flanged section
In the case of the flanged wall section shown in Figure C- 3, the factored moment capacity M r
can be determined by summing up the moments around the centroid of the wall section (point
O) as follows
 
M r  C m l w 2  x   2  s f y Ac (l w 2  d )
where
Pf   s f y Ad
AL 
0.85 m f ' m
is the area of compression zone, and its depth is
AL  b f * t  t 2
a
t

x
  
t * a 2  (b f  t ) t 2 2
2

AL
and the resultant of masonry compression stress is
C m  0.85m f 'm AL

Figure C- 3. A simplified design model for a flanged wall section.

9/1/2018 C-5
Section with boundary elements

In the case of the wall section with boundary elements shown in Figure C-4, the factored
moment capacity M r can be determined by summing up the moments around the centroid of
the wall section (point O) as follows
 
M r  C m l w 2  x   2  s f y Ac (l w 2  d )
Where
Pf   s f y Ad
AL 
0.85 m f ' m
is the area of compression zone. When the neutral axis falls within the boundary element, the
depth of compression block is
AL
a
bf
but if neutral axis falls in the wall web, the depth of the compression zone is
AL  b f  l f
a lf
t
The centroid of the masonry compression zone can be determined from the following equation:
 lf 
b f  l f  a    a  l f 
2
t 2
 2 
x
AL
and the resultant of masonry compression stress is
C m  0.85m f 'm AL

Figure C-4. A simplified design model for a wall section with boundary elements.

9/1/2018 C-6
C.1.1.2 Moment capacity for rectangular wall sections with distributed vertical
reinforcement
The previous section discussed a general case of a shear wall with both concentrated and
distributed vertical reinforcement. In low to medium-rise concrete and masonry wall structures,
the provision of distributed vertical reinforcement is often sufficient to resist the effects of
combined flexure and axial loads (see Figure C-5a)). The factored moment capacity for walls
with distributed vertical reinforcement can be determined based on the approximate equation
proposed by Cardenas and Magura (1973), which was originally developed for reinforced
concrete shear walls. The equation was derived based on the assumption that the distributed
wall reinforcement shown in Figure C-5b) can be modeled like a thin plate of length l w (equal to
the wall length), and the thickness is such that the total area Avt is the same as that provided by
distributed reinforcement along the wall length. The factored moment capacity can be
determined as follows:
 Pf  
M r  0.5 s f y Avt l w 1  1  c  (5)
  f A  l 
 s y vt  w 

where
Avt - the total area of distributed vertical reinforcement
c - neutral axis depth
 s f y Avt

m f 'm l wt
Pf

m f 'm lwt
c  

l w 2   1  1

 1  0.85 and 1  0.8

Figure C-5. Shear wall with distributed vertical reinforcement: a) vertical elevation; b) actual
cross section; c) equivalent cross-section.

9/1/2018 C-7
C.1.1.3 An approximate method to estimate the wall reinforcement
Consider the wall cross-section shown in Figure C-6a). In design practice, there is often a need
to produce a quick estimate of wall reinforcement based on the given factored loads. In this
case, the loads consist of the factored bending moment M f and the axial force Pf acting at the
centroid of the wall section (point O).

The goal of this procedure is to find the total area of wall reinforcement As . To simplify the
calculations, an assumption is made that the reinforcement yields in tension and that the
resultant force T r acts at the centroid of the wall section, that is, (see Figure C-6b)).
Tr   s f y As ( 6)
Initially, the compression zone depth a can be estimated in the range from 0.2lw to 0.3lw . The
moment resistance is usually not too sensitive to the a value as long it is relatively small. For
example, the designer could use an estimate a  0.3l w .

Figure C-6. Reinforcement estimate: a) plan view wall cross-section; b) distribution of internal
forces.
Next, compute the sum of moments of all forces around the centroid of the compression zone
(point C), as follows
M f  Pf (l w  a ) 2  Tr (l w  a ) 2  0
From the above equation it follows that
M f  Pf (l w  a ) 2
Tr  ( 7)
(l w  a ) 2
The area of reinforcement can then be determined from equation (7) as follows
As  Tr  s f y
The area of reinforcement estimated by this procedure is usually close to the required value. A
uniform reinforcement distribution over the wall length is recommended for seismic design,
since research studies have shown that shear walls with a uniform reinforcement distribution
show better seismic response in the post-cracking range. In addition, the seismic detailing
requirements for vertical reinforcement need to be followed.

9/1/2018 C-8
C.1.2 Reinforced Masonry Walls Under Out-of-Plane Seismic Loading
Masonry walls are subjected to the effects of seismic loads acting perpendicular to their surface
– this is called out-of-plane seismic loading. For design purposes, wall strips of a predefined
width are treated as beams
spanning vertically or
horizontally between lateral
supports. When the walls
span in the vertical
direction, floor and/or roof
diaphragms provide the
lateral supports.

Walls can also span


horizontally, in which case
the lateral supports need to
be provided by cross walls
or pilasters, as shown in
Figure C-7. Note that
support on four edges is
very efficient, since these
walls behave as two-way
slabs.

Figure C-7. Masonry walls under out-of-plane seismic loads: a) spanning vertically between
floor/roof diaphragms; b) spanning horizontally between pilasters.
Consider a reinforced concrete masonry wall subjected to the effects of a factored axial load
Pf and a bending moment M f , as shown in Figure C-8a). The wall is reinforced vertically, with

9/1/2018 C-9
only the reinforced cores grouted. It is assumed that the size and distribution of vertical
reinforcement are given. The notation used in Figure C-8b) is explained below:
t - overall wall thickness (taken as actual block width, e.g. 140 mm, 190 mm, etc.)
t f - face shell thickness
b - effective width of the compression zone (see Section 2.4.2 and Figure 2-19)
d - effective depth, that is, distance from the extreme compression fibre to the centroid of the
wall reinforcement; typically, the reinforcement is placed in the centre of the wall, so
d t 2
As - total area of steel reinforcement placed within the effective width b

It is assumed that the steel has yielded, that is,  s   y , and the corresponding stress in the
reinforcement is equal to the yield stress, f y . This is a reasonable assumption for low-rise
masonry buildings, since the axial load is low and the walls are expected to fail in the steel-
controlled mode. The design procedure is outlined below.

 The resultant forces in steel Tr and masonry C m can be determined as follows:


Tr   s f y As
C m  0.85 m f ' m b  a 
 The equation of equilibrium of internal forces gives (see Figure C-8d))
C m  Pf  Tr
 The depth of the compression stress block a is equal to
Cm
a ( 8)
0.85 m f ' m b
 The moment resistance can be found from the following equation
M ' r  C m ( d  a 2) (9)

9/1/2018 C-10
Figure C-8. A wall under axial load and out-of-plane bending: a) vertical section showing
factored loads; b) plan view of a wall cross-section; c) strain distribution; d) internal force
distribution.
For partially grouted wall sections (where only reinforced cores are grouted), the designer needs
to confirm that
a tf
When the above relation is correct, then the compression zone is rectangular, as shown in
Figure C-9a). Note: in solidly grouted walls, the compression zone is always rectangular!

When a  t f , the compression zone needs to be treated as a T-section and an additional


calculation is required to determine the a value. The following equations can be used to
determine the moment resistance in sections with a T-shaped compression zone:
 The resultant force in the steel Tr can be determined as follows:
Tr   s f y As

9/1/2018 C-11
 The resultant force in the masonry, C m , acts at the centroid of the compression zone and
can be determined from the equation of equilibrium of internal forces, that is,
C m  Pf  Tr
Once the compression force in the masonry is found, the area of the masonry compression
zone, Am (see Figure C-9b)), is given by
C m  0.85 m f ' m   Am
 The depth of the compression stress block a can be found from the following equation
 
Am  b  t f  a  t f  bw
where
bw = width of the grouted cell plus the adjacent webs
 The distance from the extreme compression fibre to the centroid of the compression zone a
is equal to

 2
   
b  t f 2  a  t f   t f 
atf 
2 

a  (10)
Am

Figure C-9. Masonry compression zone: a) rectangular shape; b) T-shape; c) effective width
and tributary width.

9/1/2018 C-12
 The moment resistance can be found from the following equation
M ' r  C m d  a  (11)
Note that M ' r denotes the moment capacity for a wall section of width b . It is usually more
practical to convert the M ' r value to a unit width equal to 1 metre (see Figure C-9c)), as follows
M r  M ' r 1.0 s  (12)
where
s - spacing of vertical reinforcement expressed in metres (where b  s )
M r - factored moment capacity in kNm/m.

The design of masonry walls subjected to the combined effects of axial load and bending is
often performed using P-M interaction diagrams. The axial load capacity is shown on the vertical
axis of the diagram, while the moment capacity is shown on the horizontal axis. The points on
the diagram represent the combinations of axial forces and bending moments corresponding to
the capacity of a wall cross-section. An interaction diagram is defined by the following four
distinct points and/or regions: 1) balanced point, 2) points controlled by steel yielding, 3) points
controlled by masonry compression, and 4) pure compression (zero eccentricity). A conceptual
wall interaction diagram is presented in Figure C-10.

Figure C-10. P-M interaction diagram.

1. Balanced point
At the load corresponding to the balanced point, the steel has just yielded, that is,  s   y . The
position of the neutral axis cb can be determined from the following proportion (refer to strain
diagram in Figure C-8c)):
cb 
 m
d  cb  y
or
m
cb  d ( )
m   y

9/1/2018 C-13
For f y  400 MPa and  y  0.002 it follows that
c b  0 .6 d

2. Points controlled by steel yielding


For c  c b , the steel will yield before the masonry reaches its maximum useful strain (0.003).
Since the steel is yielding, it follows that  s   y . The designer needs to assume the neutral
axis depth ( c ) value so that c  c b . The compression zone depth can then be calculated as
a   1c  0.8c (this is valid for f m  20 MPa according to S304-14 Cl.10.2.6). Combinations of
axial force and moment values corresponding to an assumed neutral axis depth can be found
from the following equations of equilibrium (see Figure C-8d)).
Pr  C m  Tr
where
Tr   s f y As (note that the stress in the steel is equal to f y since the steel is yielding)

Moment resistance depends on the shape of the masonry compression zone, that is, on
whether the section is partially or solidly grouted.
 For a solidly grouted section or a partially grouted section with the compression zone in the
face shells only:
M ' r  C m ( d  a 2)
where
C m  0.85 m f ' m b  a 
 For a partially grouted section with the compression zone extending into the grouted cells:
M ' r  C m d  a 
where
C m  0.85 m f ' m   Am

3. Points controlled by masonry compression


For c  cb , the steel will remain elastic, that is,  s   y and f s  f y , while the masonry reaches
its maximum strain of 0.003. The designer needs to assume the neutral axis depth ( c ) value so
that c  cb , and the strain in steel can then be determined from the following proportion (see
Figure C-8c)):
m s

d d c
thus
d c
s  m 
 c 
The stress in the steel can be determined from Hooke’s Law as follows
f s  E s *  s (note that steel stress f s  f y )
where E s is the modulus of elasticity for steel. The equations of equilibrium are the same as
used in part 2 above, except that
Tr   s f s As
The point corresponding to c  t 2 is considered as a special case. At that point, the strain
distribution is defined by the following values
 m  0.003 and  s  0

9/1/2018 C-14
thus
Tr  0

4. Pure compression (zero eccentricity)


In the case of pure axial compression (S304-14 Cl.10.4.1) the axial load resistance for untied
sections can be determined as follows:
Pr  0.85 m f m Ae actual axial compression resistance
and
Pr max  0.8 Pr design axial compression resistance
According to S304-14 Cl.10.4.2, when the steel bars are tied as specified in Cl.12.2, then the
steel contribution can be considered for the compression resistance. The design equation for
tied wall sections is as follows:
Pr  0.85 m f m ( Ae  As )   s f y As
and
Pr max  0.8 Pr

C.2 Wall Intersections and Flanged Shear Walls


Flanged shear wall configurations are encountered when a main shear wall intersects a cross-
wall (or transverse wall). Examples of flanged walls in masonry buildings are very common,
since the bearing wall systems often consist of walls laid in two orthogonal directions. Also, in
medium-rise wood frame apartment buildings, elevator shafts are usually of masonry
construction, and the intersecting masonry walls that form the core can be considered as
flanged walls.

C.2.1 Effective Flange Width

10.6.2

In flanged shear walls, a portion of the cross wall is considered to act as the flange, while the
main shear wall acts at the web. Depending on the cross-wall configuration, flanged shear walls
may be of I, T- or L-section. An I-section is characterized by the two end flanges, similar to that
in Figure C-11 (left), a T-section is characterized with one flanged end and one rectangular/
non-flanged end, while a L-section is characterized by one flanged end (similar to that shown in
Figure C-11 (right), and one rectangular-shaped (non-flanged) end. Design codes prescribe the
maximum effective flange width that may be considered in the shear wall design. The CSA
S304-14 requirements for overhanging flange widths for these wall sections are summarized in
Table C-1 and Figure C-11. For masonry buildings with substantial flanges the height ratio limits
will usually govern.

9/1/2018 C-15
Table C-1. Overhanging Flange Width Restrictions for T- and L- Section Walls per CSA S304-
14 Cl.10.6.2
T-sections ( bT ) L-sections ( bL ) where
bT  the smallest of: bL  the smallest of:
bactual - actual overhang/flange width
a) bactual a) bactual a w - clear distance between the
adjacent cross walls
b) a w 2 b) a w 2
t - actual flange thickness
c) 6  t c) 6  t hw - wall height
d) hw 12 d) hw 16

Figure C-11. CSA S304-14 flange width requirements.

C.2.2 Types of Intersections

According to Cl. 7.11, the effective shear transfer across the web-to-flange connection in both
unreinforced and reinforced masonry walls can be achieved through bonded or unbonded
intersections, as follows (see Figure C-12):
a) Bonded intersections – alternating courses with the units of one wall embedded at least
90 mm into the other wall (Cl.7.11.1),
b) Unbonded intersections (Cl.7.11.2) which can be achieved in the following ways:
 Mechanical connection with steel connectors (e.g. anchor straps, rods, or bolts)
at a maximum vertical spacing of 600 mm, and
 Connection with a minimum of two 3.65 mm diameter steel wires from joint
reinforcement spaced at a maximum of 400 mm vertically, or
 Fully grouted bond beam intersections with reinforcing bars spaced at 1200 mm
or less vertically.
 Steel connectors, joint reinforcing and reinforcing bars should be detailed to
develop the full yield strength on each side of the intersection.

9/1/2018 C-16
Note that S304-14 Cl.10.11.2 does not permit the use of rigid anchors (approach b)) or joint
reinforcement (approach c)) for portions of reinforced masonry shear walls in which the flanges
contain tensile steel and are subject to axial tension, but reinforced bond beams (approach d))
may be used.

Figure C-12. Masonry wall intersections: a) bonded intersections; b) mechanical connection;


c) horizontal joint reinforcement; d) horizontal reinforcing bars (bond beam reinforcement).

Seismic studies in the U.S. under the TCCMAR research program resulted in recommendations
related to horizontal reinforcement at the web-to-flange intersections (Wallace, Klingner, and
Schuller, 1998). To ensure the effective shear transfer, horizontal reinforcement in bond beams
needs to be continued from one wall into other, for a distance of 600 mm (2 feet) or 40 bar
diameters, whichever is greater. The grout must be continued across the intersection by
removing the face shells of the masonry units in one of the walls, as illustrated in Figure C-13.
Note that TMS 402/602-16 requires that bond beams in ductile walls be provided at a vertical
spacing of 1200 mm (4 feet).

9/1/2018 C-17
Figure C-13. Horizontal reinforcement at the web-to-flange intersection: TCCMAR
recommendations.

C.2.3 Shear Resistance at the Intersections

7.11

Vertical shear resistance of the intersections must be checked by one of the following methods:
 For bonded intersections, vertical shear at the intersection shall not exceed the out-of-plane
masonry shear resistance (Cl.7.10.2).
 For flanged sections with the mechanical steel connectors (Figure C-12 approach b), the
connectors must be capable of resisting the vertical shear at the intersection. The connector
resistance should be determined according to CSA A370-14.
 For flanged sections with the horizontal reinforcement (approaches c and d), the
reinforcement must be capable of resisting the vertical shear at the intersection.

Vertical shear resistance for bonded wall intersections

7.11.1

The factored vertical shear resistance at bonded intersections should not exceed the factored
shear resistance of the masonry taken as
Vr  0.16m f m Ae
where Ae is effective mortared area of the bed joint for hollow and partially grouted walls. For
fully grouted walls Ae is gross cross-sectional area.
Minimum horizontal reinforcement shall be provided across the vertical intersection. This
reinforcement shall be equivalent in area to at least two 3.65 mm diameter steel wires spaced
400 mm vertically.

9/1/2018 C-18
Vertical shear resistance for unbonded wall intersections

7.11.2

Where wall intersections are not bonded in accordance with Cl.7.11.1, or where additional
capacity is required, the factored shear resistance of the web-to-flange joint shall be based on
the shear friction resistance taken as
Vr   m C h
where
 = 1.0 coefficient of friction for the web-to-flange joint
C h = compressive force in the masonry acting normal to the head joint, normally taken as the
factored tensile force at yield of the horizontal reinforcement that crosses the vertical section.
The reinforcement must be detailed to enable it to develop its yield strength on both sides of the
vertical masonry joint, which may be hard to achieve in practice.

Commentary

For flanged walls with horizontal reinforcement, resistance to vertical shear sliding is provided
by the frictional forces between the sliding surfaces, that is, the web and the flange of the wall.
The shear friction resistance Vr is proportional to the coefficient of friction  , and the clamping
force C h acting perpendicular to the joint of height h (see Figure C-14a)).

C h is equal to the sum of the tensile yield forces developed in reinforcement of area As spaced
at the distance s , that is,
Ch   s f y As h s
In case of a flanged shear wall with openings, shear friction resistance Vr is provided by wall
segments between the openings, as shown in Figure C-14b).

Reinforcement providing the shear friction resistance should be distributed uniformly across the
joint. The bars should be long enough so that their yield strength can be developed on both
sides of the vertical joint, as shown in Figure C-15b).

Cl.7.11.2 lists three approaches (a, b, and c) that can be used to ensure shear transfer at the
web-to-flange interface for unbonded masonry. The U.S. masonry design standard TMS
402/602-16 prescribes intersecting bond beams in intersecting walls at maximum spacing of
1200 mm (4 ft) on centre. The bond beam reinforcement area shall not be less than 200 mm2
per metre of wall height (0.1 in2/ft), and the reinforcement shall be detailed to develop the full
yield stress at the intersection.

9/1/2018 C-19
Figure C-14. Shear friction resistance at the web-to-flange intersection: a) resistance provided
by the reinforcement; b) flanged shear wall with openings.

When the shear resistance of the web-to-flange interface relies on masonry only (see Figure C-
15a)), the horizontal shear stress v f , due to shear force V f , can be given by:
Vf
vf 
telw
where
t e - effective web width
l w - wall length
The designer should also find the vertical shear stress caused by the resultant compression
force Pfb :
Pfb
vf 
b w * hw
The larger of these two values governs. The factored shear stress should be less than the
factored masonry shear resistance, vm , as follows

9/1/2018 C-20
v f  vm

where
vm  0.16m f m
If the above condition is not satisfied, horizontal reinforcement needs to be provided
(see Figure C-15b)), and the following shear resistance check should be used
vr  vm  vs
and
v f  vr
where v s is the factored shear resistance provided by the steel reinforcement, which can be
determined as follows:
s As f y
vs 
s  te
where As is area of horizontal steel reinforcement crossing the web-to-flange intersection at the
spacing s .

Note that the reinforcement that crosses the vertical section has to be detailed to develop yield
strength on both sides of the vertical masonry joint (see Figure C-15b)).

Figure C-15. Shear resistance of the web-to-flange interface: a) bonded masonry intersection;
b) horizontal reinforcement at the intersection.

9/1/2018 C-21
C.3 Wall Stiffness Calculations
The determination of wall stiffness is one of the key topics in the seismic design of masonry
walls. Although this topic has been covered in other references (e.g. Drysdale and Hamid, 2006,
and Hatzinikolas, Korany and Brzev 2015), a few key concepts are discussed in this section.
Section C.3.2 derives expressions for the in-plane lateral stiffness of walls under the assumption
that the walls are uncracked. For seismic analysis it is expected that the walls will be pushed
into the nonlinear range, and so cracking will occur and the reinforcement will yield. The
stiffness to be used in seismic analysis should not be the linear elastic (uncracked) stiffness but
some effective stiffness that reflects the effect of cracking up to the yield capacity of the wall.
Section C.3.5 gives some suggestions for the effective stiffness of shear walls responding in
shear-dominant and flexure-dominant modes.

C.3.1 Lateral Load Distribution


The distribution of lateral seismic loads to individual walls can be performed once the storey
shear forces have been determined from the seismic analysis. The flexibility of floor and/or roof
diaphragms is one of the key factors influencing the load distribution (for more details, see
Example 3 in Chapter 3). In the case of a flexible diaphragm, the lateral storey forces are
usually distributed to the individual walls based on the tributary area. In the case of a rigid
diaphragm, these forces are distributed in proportion to the stiffness of each wall. In calculating
the wall forces, torsional effects must be considered, as discussed in Section 1.11. The
distribution of lateral loads (without torsional effects) in a single-storey building with a rigid
diaphragm is shown in Figure C-16.

Figure C-16. Distribution of lateral loads to individual walls.


Wall stiffness is usually determined from the elastic analysis, and depends on wall height/length
aspect ratio, thickness, mechanical properties, extent of cracking, size and location of openings,
etc.

9/1/2018 C-22
C.3.2 Wall Stiffness: Cantilever and Fixed-End Model
Wall stiffness depends on the end support conditions, that is, whether a wall or pier is fixed or
free to move and/or rotate at its ends. Two models for wall stiffness include the cantilever model
and the fixed-end model, as shown in Figure C-17. In the cantilever model, the wall is free to
rotate and move at the top in the horizontal direction – this is usually an appropriate model for
the walls in a single-storey masonry building.

The stiffness can be defined as the lateral force required to produce a unit displacement, but it
is determined by taking the inverse of the combined flexural and shear displacements produced
by a unit load. It should be noted that flexural displacements will govern for walls with an aspect
ratio of 2 or higher. For example, the contribution of shear deformation in a wall with a
height/length aspect ratio of 2.0, is 16% for the cantilever model and 43% for the fixed-end
model. The stiffness equations presented in this section take into account both shear and
flexural deformations.

The stiffness of a cantilever wall or a pier can be determined from the following equation (see
Figure C-17 a)):
K 1
 ( 13)
Em * t  h   h 
2

  4   3
 l w    l w  
The stiffness of a wall or a pier with the fixed ends can be determined from the following
equation (see Figure C-17 b)):
K 1
 ( 14)
Em * t  h   h 
2

     3
 lw   l w  
where
h - wall height (cantilever model) or clear pier height (fixed-end model)
l w - wall or pier length
E m  850 f m modulus of elasticity for masonry
The following assumptions have been taken in deriving the above equations:
G m  0 .4 E m modulus of rigidity for masonry (shear modulus)
3
te * lw
I uncracked wall moment of inertia
12
5 * te * lw
Av  shear area (applies to rectangular wall sections only)
6
where t e = effective wall thickness.

9/1/2018 C-23
Figure C-17. Wall stiffness models: a) cantilever model, and b) fixed-end model.
The wall stiffnesses for both models for a range of height/length aspect ratios are presented in
Table D-3. Note that the derivation of stiffness equations has been omitted since it can be found
in other references (see Hatzinikolas, Korany and Brzev 2015).

C.3.3 Approximate Method for Force Distribution in Masonry Shear Walls


In most real-life design applications, walls are perforated with openings (doors and windows).
The seismic shear force in a perforated wall can be distributed to the piers in proportion to their
stiffnesses. This approach is feasible when the openings are very large and the stiffness of lintel
beams is small relative to the pier stiffnesses, or if the lintel beam is very stiff so that connected
piers act as fixed-ended walls. Figure C-18 illustrates the distribution of the wall shear force V
to individual piers in direct proportion to their stiffness. Note that, according to this model, the
wall shear force is equal to the sum of shear forces in the piers, that is,
V   Vi
where
Vi  K i *  i force in the pier i
Thus
V   (K i *  i )
If the floor diaphragm is considered to be rigid, it can be assumed that the lateral displacement
in all piers is equal to  , that is,
 A   B  C  
and so
V  ( K i ) * 
Thus
V

 Ki
where
K   Ki
denotes the overall wall stiffness for the system.

9/1/2018 C-24
Therefore, the force in each pier is proportional to its stiffness relative to the sum of all pier
stiffnesses within the wall, as follows
V Ki
Vi  K i *  i  K i * V *
 Ki  Ki
This means that stiffer piers are going to attract a larger portion of the overall shear force. This
can be explained by the fact that a larger fraction of the total lateral force is required to produce
the same deflection in a stiffer wall as in a more flexible one.

Figure C-18. Shear force distribution in a wall with a rigid diaphragm: a) wall in the deformed
shape: b) pier forces.
An approximate approach for determining the stiffness of a solid shear wall in a multi-storey
building is to consider the structure as an equivalent single-storey structure, as shown in Figure
C-19. The entire shear force is applied at the effective height, he , defined as the height at which
the shear force V f must be applied to produce the base moment M f , that is,
Mf
he 
Vf
The wall stiffness is found to be equal to the reciprocal of the deflection at the effective height
 e , as follows

9/1/2018 C-25
1
K
e
This model, although not strictly correct, can be used to determine the elastic distribution of the
torsional forces as well as the displacements, as illustrated in Example 2 in Chapter 4.

Figure C-19. Vertical combination of wall segments with different stiffness properties.
Several different elastic analysis approaches can be used to determine the stiffness of a wall
with openings. A simplified approach suitable for the stiffness calculation of a perforated wall in
a single-storey building can be explained with the help of an example of the wall X 1 shown in
Figure C-20 (see also Example 3 in Chapter 3). For a unit load applied at the top, the wall
stiffness calculation involves the following steps:
 First, calculate the deflection at the top for a cantilever wall, considering the wall to be solid
(  solid ).
 Next, calculate the deflection for the strip containing openings (  strip ), considering the full
wall length (i.e. ignore openings).
 Finally, calculate the deflection for the piers A, B, C, and D (  ABCD ) assuming that all piers
have the same deflection.
Note that the deflections for individual components are calculated as the inverse of their
stiffness values, and that the pier stiffnesses are determined assuming either the cantilever or
fixed-end models. In most cases, the use of the cantilever model is more appropriate.

Figure C-20. An example of a perforated wall.


The overall wall deflection can be determined by combining the deflections for these
components, as follows:
   solid   strip   ABCD

9/1/2018 C-26
Note that the strip deflection is subtracted from the solid wall deflections - this removes the
entire portion of the wall containing all the openings, which is then replaced by the four
segments.

Finally, the wall stiffness is equal to the reciprocal of the deflection, as follows
1
K

C.3.4 Advanced Design Approaches for Reinforced Masonry Shear Walls


with Openings
The approximate approach based on elastic analysis presented in Section C.3.3 is appropriate
for determining the lateral force distribution in masonry walls. However, that method is not
adequate for predicting the strengths in perforated reinforced masonry shear walls (walls with
openings). Openings in a masonry shear wall alter its behaviour and add complexity to its
analysis and design. When the openings are relatively small, their effect can be ignored,
however in most walls the openings need to be considered. The following two design
approaches can be used to design walls with openings:
1) Plastic analysis method, and
2) Strut-and-tie method.
These two approaches have been evaluated by experimental studies and have shown very
good agreement with the experimental results (Voon, 2007; Elshafie et al., 2002; Leiva and
Klingner, 1994). The key concepts will be outlined in this section.

C.3.4.1 Plastic analysis method


The plastic analysis method, also known as limit analysis, can be used to determine the ultimate
load-resisting capacity for statically indeterminate structures. A masonry wall with an opening as
shown in Figure C-21a) can be modeled as a frame (see Figure C-21b)). The model is
subjected to an increasing load until the flexural capacity of a specific section is reached and a
plastic hinge is formed at that location. (The plastic hinge is a region in the member that is
assumed to be able to undergo an infinite amount of deformation, and can therefore be treated
as a hinge for further analysis.) With further load increases, plastic hinges will be formed at
other sections as their flexural capacity is reached. This process continues until the system
becomes statically determinate, at which point the formation of one more plastic hinge will result
in a collapse under any additional load. This is called a collapse mechanism, and an example is
shown in Figure C-21c). There is usually more than one possible collapse mechanism for a
statically indeterminate structure, and the mechanism that gives the lowest capacity is closest to
the ultimate capacity, as this is an upper bound method.

For specific application to perforated masonry walls, the wall is idealized as an equivalent
frame, where piers are modeled as fixed at the base, and either pinned or fixed at the top, while
lintels are modeled as fixed at the ends. A failure state is reached when plastic hinges form at
the member ends, and the collapse mechanism forms. The sequence of plastic hinge formation
depends on the relative strength and stiffness of the elements. In this approach, structural
members must be designed to behave mainly in a flexural mode, while a shear failure is
avoided by applying the capacity design approach.

9/1/2018 C-27
Figure C-21. An example of a plastic collapse mechanism for a frame system: a) perforated
masonry wall; b) frame model; c) plastic collapse mechanism.
The following two mechanisms are considered appropriate for the plastic analysis of reinforced
masonry walls with openings, as shown in Figure C-22 (Leiva and Klingner, 1994; Leiva et al.
1990):

b) pier mechanism, and

c) coupled wall mechanism.

a)

b) c)
Figure C-22. Plastic analysis models for perforated walls: a) actual wall; b) pier model;
c) coupled wall model (Leiva and Klingner, 1994, reproduced by permission of The
Masonry Society).

9/1/2018 C-28
A pier mechanism is a collapse mechanism with flexural hinges at the tops and bottoms of the
piers. A pier-based design philosophy visualizes a perforated wall as a ductile frame. Horizontal
reinforcement above and below the openings is needed to transfer the pier shears into the rest
of the wall. A drawback of the pier mechanism is that the formation of plastic hinges at the top
and bottom of all piers at a story level can lead to significant damage to the piers, which are the
main vertical load-carrying elements.

A coupled wall mechanism is a collapse mechanism in which flexural hinges are formed at the
base of the wall and at the ends of the coupling lintels. A perforated wall is modeled as a series
of ductile coupled walls; this concept is similar to that used for seismic design of reinforced
concrete shear walls. The vertical reinforcement in each pier must be designed so that the
flexural capacity of the piers exceeds the flexural capacity of the coupling beams. To achieve
this, additional longitudinal reinforcement is placed in the piers, but cut off before it reaches the
wall base. The shear reinforcement in the coupling beams is designed based on the flexural and
shear capacity of the piers. Since masonry walls are usually long in plan, the formation of plastic
hinges at their bases produces large strains in the wall longitudinal reinforcement. Plastic hinges
must have adequate rotational capacity to allow the complete mechanism to form; this can be
achieved in wall structures with low axial load. To ensure the successful application of the
plastic analysis method, the wall reinforcement must be detailed to develop the necessary
strength and inelastic deformation capacity.

Figure C-23 shows a simple single-storey wall that is analyzed for the two mechanisms.
Ultimate shear forces corresponding to the pier and coupled wall mechanisms can be
determined from the equations of equilibrium assuming that the moments at the plastic hinge
locations are known. These equations are summarized in Figure C-23 (Elshafaie et al., 2002).

The plastic analysis method has a few advantages: stiffness calculations are not required, and
the designer can choose the failure mechanism, which ensures a desirable ductile response.
The designer needs to have a general background in plastic analysis, which is covered in
several references, e.g. Bruneau, Uang, and Whittaker (1998) and Ferguson, Breen, and Jirsa
(1988). This method is also used for the seismic analysis of concrete and steel structures, and
is referred to as nonlinear static analysis or pushover analysis.

9/1/2018 C-29
a)

b)
Figure C-23. Ultimate wall forces according to the plastic analysis method: a) pier mechanism;
b) coupled wall mechanism (Elshafaie et al., 2002, reproduced by permission of the Masonry
Society).

C.3.4.2 Strut-and-Tie Method


The strut-and-tie method essentially follows the truss analogy approach used for shear design
of concrete and masonry structures. Pin-connected trusses consist of steel tension members
(ties), and masonry compression members (struts). The masonry compression struts develop
between parallel inclined cracks in the regions of high shear. The essential feature of this
approach is that the designer needs to find a system of internal forces that is in equilibrium with
the externally applied loads and support conditions. A further essential feature is that the
designer must ensure that the steel and masonry tie members provided adequately resist the
forces obtained from the truss analysis.

The design of tension ties is particularly important. If a ductile response is to be assured, the
designer should choose particular tension chords in which yielding can best be accommodated.
Other ties can be designed so that no yielding will occur by using the capacity design approach.
The magnitudes of the forces in critical tension ties can be determined from statics,
corresponding to the overturning moment capacity of the wall using the nominal material
properties (rather than the factored ones). The remaining forces are then determined from the
equilibrium of nodes (conventional truss analysis). Compression forces developed in masonry
struts are usually small due to the small compression strains and do not govern the design.

9/1/2018 C-30
Careful detailing of the wall reinforcement is necessary to ensure that the actual structural
response will correspond to that predicted by the analytical model.

The designer needs to use judgement to simplify the force paths that are chosen to represent
the real structure – these differ considerably depending on individual judgement.

An example of a strut-and-tie model for a two-storey perforated masonry wall subjected to


seismic lateral load is shown in Figure C-24 (note that gravity load also needs to be considered
in the analysis, however it is omitted from the figure). It can be seen that two different models
are required to account for the alternate direction of seismic load. The examples show the
seismic load being applied as a compressive load to the building; however, these loads should
be applied to the floor levels, depending on the diaphragm-to-wall connection. The designated
tie members in one model will become struts in the other model (when the seismic load changes
direction). An advantage of the reversible nature of seismic forces is that a significant fraction of
the inelastic tensile strains imposed on the end strut members is recoverable due to force
reversal, thereby providing hysteretic energy dissipation. A detailed solution for this example is
presented in the User’s Guide by NZCMA (2004).

Figure C-24. Strut-and-tie models for a masonry wall corresponding to different directions of
seismic loading (NZCMA, 2004, reproduced by the permission of the New Zealand Concrete
Masonry Association Inc.).
Strut-and-tie models are used for the design of masonry walls in New Zealand, and this
approach is explained in more detail by Paulay and Priestley (1992). The New Zealand Masonry
Standard NZS 4230:2004 (SANZ, 2004) recommends the use of strut-and-tie models for the
design of perforated reinforced masonry shear walls. In Canada, strut-and-tie models are used
to design discontinuous regions of reinforced concrete structures according to the Standard
CSA A23.3-04 Design of Concrete Structures. The design concepts and applications of strut-
and-tie models for concrete structures in Canada are covered by McGregor and Bartlett (2000).

9/1/2018 C-31
C.3.5 The Effect of Cracking on Wall Stiffness
The behaviour of masonry walls under seismic load conditions is rather complex and depends
on the failure mechanism (shear-dominant or flexure-dominant), as discussed in Section 2.3.1.
Figure C-25 shows the hysteretic response of shear-dominant and flexure-dominant walls. The
effective stiffness discussed in this section reflects the secant stiffness up to first crack in brittle
shear-dominant walls, and the stiffness for an elastic-perfectly-plastic model that would
approximate the strength envelope of the hysteretic plot in ductile flexure-dominant walls.

For the shear-dominant mechanism, the response is initially elastic until cracking takes place, at
which point there is a substantial drop in stiffness. This is particularly pronounced after the
development of diagonal shear cracks. After a few major cracks develop, the load resistance is
taken over by the diagonal strut mechanism, and the shear stiffness can be estimated by an
appropriate strut model. However, the stiffness drops significantly shortly after the strut
mechanism is formed and can be considered to be zero for most practical purposes (see Figure
C-25b)). It is expected that an increase in the quantity of vertical and horizontal steel and/or the
magnitude of axial compressive stress causes a reduced crack size and an increase in the
shear stiffness (Shing et al., 1990).

Figure C-25. Cracking pattern and load-displacement curves for damaged masonry wall
specimens tested by Shing et al. (1990, 1991): a) flexure-dominant response, and b) shear-
dominant response (Kingsley, Shing, and Gangel, 2014).

9/1/2018 C-32
For the flexure-dominant mechanism, a drop in the stiffness immediately after the onset of
cracking is not very significant. As can be seen from Figure C-25a), the stiffness drops after the
yielding of vertical reinforcement takes place, and continues to drop with increasing inelastic
lateral deformations (this depends on the ductility capacity of the wall under consideration). The
specimen for which the results are shown in Figure C-25a) showed yielding of vertical
reinforcement and compressive crushing of masonry at the wall toes (Shing et al., 1989).

Note that the height of wall test specimens shown in Figure C-23 was 1.8 m (6 feet), thus a
2.5% drift ratio permitted by the NBC 2015 for regular buildings corresponds to 45 mm (1.8 inch
) displacement. It can be seen that the displacements and drifts in these specimens are very
low, particularly so for the shear-dominant specimen shown in Figure C-25b).

Evidence from studies that focus on quantifying the changes in in-plane wall stiffness under
increasing lateral loading are limited, so CSA S304-14 and other masonry codes do not provide
guidance related to this issue. Shing et al. (1990) tested a series of 22 cantilever block masonry
wall specimens that were laterally loaded at the top, with a height/length aspect ratio of 1.0.
Based on the experimental test data, they have recommended the following empirical equation
for the lateral stiffness of a wall with a shear-dominant response
K e  (0.2  0.1073 f c ) K shear  K el ( 15)
where
Em * te
K shear  is the shear stiffness of a wall/pier
h
3 *  
 lw 
h = wall height
l w = wall length
t e = effective wall thickness
f c = axial compressive stress (MPa)
The above equation is based on the force/displacement measurements taken just after the first
diagonal crack developed, in specimens with a height/length ratio of 1.0. For seismic
applications where the walls are expected to yield in flexure before failing in shear, and the
lateral stiffness is used to estimate the fundamental period of the structure and to determine the
seismic displacements, it is more appropriate to determine the effective stiffness from a cracked
section analysis at first yield of the tension reinforcement.

A study by Priestley and Hart (1989), based on the cracked transformed section stiffness at first
yield of the tension reinforcement, recommends that the effective moment of inertia, I e , of a wall
can be approximated by:
100 Pf
Ie  (  )I g ( 16)
fy f m Ae
where
f y = steel yield strength (MPa)
Pf = factored axial load
Ae = effective cross-sectional area for the wall
f m = masonry compressive strength, and

9/1/2018 C-33
t e * l w3
Ig  is the gross moment of inertia of the wall.
12
Note that the first term in the bracket, 100 f y , is equal to 0.25 for f y  400 MPa (Grade 400
steel). The second term is a ratio of axial compressive stress in the wall, equal to Pf Ae , and
the masonry compressive strength, f m .

The above relation is based solely on consideration of flexural stiffness, and is a best fit
relationship for several different values of height/length ratio ( h l w ), steel strength, vertical
reinforcement ratio and axial load. Other considerations are whether the vertical reinforcement
is uniformly distributed across the wall length or concentrated at the ends, and the effect of
tension stiffening. The vertical reinforcement ratio is not included in the above expression, and
as a result, the wall stiffness is overestimated for lightly reinforced walls and underestimated for
heavily reinforced walls.

If it is assumed that wall cracking causes the same proportional decrease in the effective shear
area as it does for the moment of inertia, then the stiffnesses can be combined to give the
following equation for the reduced wall stiffness, K ce ,
100 Pf
K ce  (  )K c (17)
fy f m Ae
where
Em * te
Kc 
 h   h  
2

  4   3
 l w    l w  

is the combined stiffness of an uncracked cantilever wall or pier, considering both the flexural
and shear deformation components (refer to Section C.3.2 for the wall stiffness equations).

The terms in the large right-hand bracket of the K c equation give the comparative value of
flexural deformation to shear deformation. At a h l w ratio of 1.0, flexure contributes 4/7 of the
total deformation and shear 3/7, while at a h l w ratio of 0.5, shear contributes ¾ of the total
defection.

The Priestley and Hart equation was obtained using experimental data related to cantilever wall
specimens, however it may also be used for fixed-end walls. The stiffness equation for these
walls, K fe , is the same as for the cantilever walls, that is,
100 Pf
K fe  (  )K f (18)
fy f m Ae
where
Em * te
Kf  is the stiffness of an uncracked fixed-end wall or a pier
 h   h  
2

     3
 l w   l w  

A comparison of the proposed equations for a masonry block wall under axial compressive
stress is presented in Figure C-26. The following values were used in the calculations:
f y  400 MPa, Pf Ae = 1 MPa, and f m = 10 MPa.

9/1/2018 C-34
Note that the Shing equation is only shown for h l w up to 1.5 as it is based entirely on shear
deformation. Since the Shing equation represents stiffness at first diagonal cracking, it is
expected to give higher stiffness values than the Priestley-Hart equation. Use of the Priestley-
Hart stiffness equation is recommended since it is valid for all h l w ratios.

The elastic uncracked stiffness could be used to distribute lateral seismic load to individual walls
and piers, but the reduced cracked stiffness should be used for period estimation and deflection
calculations.

The wall design deflections can be found from the following equation:
Rd * Ro
 design   el *
IE
where
 el = elastic deflections calculated using the reduced wall stiffness ( K ce or K fe ) and the
factored design forces, and
R d * Ro
= deflection multiplier to account for the effects of ductility, overstrength, and the
IE
building importance factor (see Section 1.13)
0.25

0.20 Shing

Priestley

0.15
K/(E*t)

0.10

0.05

0.00
0.5 1 1.5 2 2.5 3
h/l

Figure C-26. A comparison of the stiffness values obtained using the Shing and Priestley-Hart
equations.

9/1/2018 C-35
TABLE OF CONTENTS – APPENDIX D

D DESIGN AIDS D-2

Table D-1. Properties of Concrete Masonry Walls (per metre or foot length) D-2
Table D-2. c l w ratio, f y = 400 MPa D-3
Figure D-1. c l w ratio, f y = 400 MPa D-4
Table D-3. Wall Stiffness Values K E m * t  D-5

9/1/2018 D-1
D Design Aids
Table D-1. Properties of Concrete Masonry Walls (per metre or foot length) 1

1Source: Masonry Technical Manual (MIBC, 2017, reproduced by permission of the Masonry Institute of
BC)

9/1/2018 D-2
Table D-2. c l w ratio, f y = 400 MPa

Input parameters: Units:

Avt Pf (kN)
 vflex 
t * lw l w , t (mm)
566.7 *  vflex Avt (mm2)

f 'm f ' m (MPa)
1667 * Pf

f 'm l wt

9/1/2018 D-3
Figure D-1: c l w ratio, f y = 400 MPa

9/1/2018 D-4
Table D-3. Wall Stiffness Values K E m * t 

Cantilever model:

K 1

Em * t  h   h 
2

  4   3
 l w    l w  

Fixed both ends:

K 1

Em * t  h   h 
2

     3
 lw   l w  

E m  850 f m Modulus of
elasticity
G  0 .4 E m Modulus of
rigidity (shear modulus)
Av  5A 6 Shear area

9/1/2018 D-5
E Notation

a max = maximum acceleration


a = depth of the compression zone (equivalent rectangular stress block)
a w = clear distance between the adjacent cross walls
Ab = area of reinforcement bar
Ac = area of concentrated reinforcement at each end of the wall
Ach = cross-sectional area of core of the boundary element
Ad = area of distributed reinforcement along the wall length
Ae = effective cross-sectional area of masonry
Ag = gross cross-sectional area of masonry

AL = area of the compression zone (flanged wall section)


Ar = response amplification factor to account for the type of attachment of equipment or veneer ties
As = area of steel reinforcement
Ash = total area of rectangular hoop reinforcement (buckling prevention ties) in each horizontal direction
of the boundary element

Auc = uncracked area of the cross-section


Av = area of horizontal wall reinforcement
Avt = total area of the distributed vertical reinforcement
Av = shear area of the wall section
Ax = amplification factor at level x to account for variation of response with the height of the building
(veneer tie design)
b = effective width of the compression zone
bactual = actual flange width
bc = critical wall thickness
bT = overhanging flange width
bw = overall web width (shear design)
B = torsional sensitivity factor
c = neutral axis depth (distance from the extreme compression fibre to the neutral axis)
C = compressive force in the masonry acting normal to the sliding plane

9/1/2018 E-1
C m = the resultant compression force in masonry
C h = compressive force in the masonry acting normal to the head joint
C p = seismic coefficient for a nonstructural component (veneer tie design)
d = effective depth (distance from the extreme compression fibre to centroid of tension reinforcement)
d v = effective wall depth for shear calculations
d  = distance from the extreme compression fibre to the centroid of the concentrated compression
reinforcement

Dnx = plan dimension of the building at level x perpendicular to the direction of seismic loading being
considered
e = load eccentricity
ea = accidental torsional eccentricity
e x = torsional eccentricity (distance measured perpendicular to the direction of earthquake loading
between the centre of mass and the centre of rigidity at the level being considered)
E f = modulus of elasticity of the frame material (infill walls)

E m = modulus of elasticity of masonry


f t = flexural tensile strength of masonry (see Table 5 of CSA S304-14)
f m = compressive strength of masonry normal to bed joints at 28 days (see Table 4 of CSA S304-14)
f y = yield strength of reinforcement

f yh = specified yield strength of hoop reinforcement in a boundary element


F = force
F (T ) = site coefficient (NBC 2015 Cl.4.1.8.4)
Ft = a portion of the base shear V applied at the top of the building
Fel = elastic force
Fs = factored tensile force at yield of horizontal reinforcement
Fa = acceleration-based site coefficient
Fv = velocity-based site coefficient
Fx = seismic force applied to level x
Fy = yield force
G = modulus of rigidity for masonry (shear modulus)

9/1/2018 E-2
h = unsupported wall height/height of the infill wall
hc = dimension of core of rectangular section measured perpendicular to the direction of the hoop bars
(boundary elements)

hn = building height
hp = extent of the plastic hinge region above the critical section of the shear wall (previously l p )

hs = storey height
hw = total wall height

hx = height from the base of the structure up to the level x


I b = moment of inertia of the beam
I c = moment of inertia of the column
I E = earthquake importance factor of the structure
J = numerical reduction coefficient for base overturning moment
k = effective length factor for compression member
kn = factor accounting for the effectiveness of transverse reinforcement in a boundary element
k p1 = factor accounting for the compressive strain level in a boundary element
K = stiffness
l = length of the infill wall
l d = length of the diagonal (infill wall)
l s = design length of the diagonal strut (infill wall)
l w = wall length
Ln = clear vertical distance between lines of effective horizontal support or clear horizontal distance
between lines of effective vertical support
M = mass
M f = factored bending moment

M r = factored moment resistance


M n = nominal moment resistance
M p = probable moment resistance

M v = factor to account for higher mode effect on base shear

9/1/2018 E-3
nl = total number of longitudinal bars in the boundary element cross-section that are laterally supported
by the corner of hoops or by hooks of seismic cross-ties
N = axial load arising from bending in coupling beams or piers
p f = distributed axial stress

PGAref = reference Peak Ground Acceleration (PGA) for determining F (T )

Pd = axial compressive load on the section under consideration


Pcr = critical axial compressive load
PDL = dead load
Pfb = the resultant compression force (flanged walls)

Pr = factored axial load resistance


P1 = compressive force in the unreinforced masonry acting normal to the sliding plane
Ph = horizontal component of the diagonal strut compression resistance (infill walls)
Pv = the vertical component of the diagonal strut compression resistance (infill walls)
Pult = ultimate tie strength
Rd = ductility-related force modification factor
Ro = overstrength-related force modification factor
R p = element or component response modification factor (veneer tie design)
s = reinforcement spacing
S T  = design spectral acceleration
S a (T ) = 5% damped spectral response acceleration
S e = section modulus of effective wall cross-sectional area
S p = horizontal force factor for part or portion of a building and its anchorage (veneer tie design)
t = overall wall thickness
t e = effective wall thickness
t f = face shell thickness
T = fundamental period of vibration of the building
Tx = torsional moment at level x
T r = the resultant force in steel reinforcement
T y = factored tensile force at yield of the vertical reinforcement

9/1/2018 E-4
v f = distributed shear stress

vm = masonry shear strength


v max = maximum velocity
V = lateral earthquake design force at the base of the structure
Ve = lateral earthquake elastic force at the base of the structure
V f = factored shear force

V fr = shear flow resistance

Vnb = the resultant shear force corresponding to the development of nominal moment resistance M n at
the base of the wall

Vm = masonry shear resistance


Vr = factored shear resistance
Vs = average shear wave velocity in the top 30 m of soil or rock
Vs = factored shear resistance of steel reinforcement
w = diagonal strut width (infill walls)
we = effective diagonal strut width (infill walls)
W = seismic weight, equal to the dead weight plus some portion of live load that would move laterally
with the structure
W p = weight of a part or a portion of a structure (veneer tie design)

W x = a portion of seismic weight W that is assigned to level x


h = vertical contact length between the frame and the diagonal strut (infill walls)

L = horizontal contact length between the frame and the diagonal strut (infill walls)

 = damping ratio

d = ratio of the factored dead load moment to the total factored moment

1 = ratio of depth of rectangular compression block to depth of the neutral axis

 g = factor to account for partially grouted or ungrouted walls that are constructed of hollow or semi-solid
units

 max = the maximum storey displacement at level x at one of the extreme corners in the direction of

earthquake

9/1/2018 E-5
 ave = the average storey displacement determined by averaging the maximum and minimum

displacements of the storey at level x


 = lateral displacement
 p = plastic displacement

 y = displacement at the onset of yielding

 el = elastic displacement
 max = maximum displacement
 u = inelastic (plastic) displacement
m = the maximum compressive strain in masonry

s = strain in steel reinforcement

y = yield strain in steel reinforcement

 = factor used to account for direction of compressive stress in a masonry member relative to the

direction used for determination of f m


 = curvature

u = ultimate curvature

y = yield curvature corresponds to the onset of yielding

 er = resistance factor for member stiffness

m = resistance factor for masonry

s = resistance factor for steel reinforcement

 = resistance factor

h = horizontal reinforcement ratio

s = volumetric ratio of circular hoop reinforcement for buckling prevention ties

v = vertical reinforcement ratio

 = coefficient of friction

 = displacement ductility ratio

 = curvature ductility ratio

 = displacement ductility ratio

 = angle of diagonal strut measured from the horizontal

9/1/2018 E-6
 e = elastic rotation
ic = inelastic rotational capacity
id = inelastic rotational demand
p = plastic rotation

 = natural frequency

9/1/2018 E-7

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy