You are on page 1of 9

APPROXIMATE SEISMIC LATERAL DEFORMATION DEMANDS IN

MULTISTORY BUILDINGS
By Eduardo Miranda1

ABSTRACT: An approximate method to estimate the maximum lateral deformation demands in multistory
buildings responding primarily in the fundamental mode when subjected to earthquake ground motions is pre-
sented. This method permits a rapid estimation of the maximum roof displacement and of the maximum interstory
drift for a given acceleration time history or for a given displacement response spectrum. A multistory building
is modeled as an equivalent continuum structure consisting of a combination of a flexural cantilever beam and
a shear cantilever beam. The simplified model is used to investigate the ratio of the spectral displacement to the
roof displacement and the ratio of the maximum interstory drift ratio to the roof drift ratio. The effect of the
Downloaded from ascelibrary.org by University of Leeds on 09/17/13. Copyright ASCE. For personal use only; all rights reserved.

distribution of lateral forces along the height of the building and of the ratio of overall flexural and shear
deformations is examined. Lateral deformation demands of a 10-story steel building computed with the simplified
method when subjected to various earthquake ground motions are compared with those computed using step-
by-step time history analyses. It is shown that the method provides good approximations, which are useful for
the preliminary design of new buildings or for a rapid evaluation of existing buildings.

INTRODUCTION ground motions; and (2) to compare the results of the proposed
method with those computed with detailed step-by-step time
Both structural and nonstructural damage sustained during history analyses. The approximate method is intended to be
earthquake ground motions are produced primarily by lateral used during the preliminary design of new buildings and for
displacements. Thus, adequate damage control can be achieved a rapid evaluation of existing buildings and is not intended to
if lateral deformations are controlled by providing enough lat- be a substitute of more detailed analyses, which are appropri-
eral stiffness, lateral strength, and energy dissipation capacity ate during the final evaluation of the proposed design of a new
to a structure. However, current building codes are based on building or during the detailed evaluation of existing build-
lateral forces and give a secondary importance to lateral dis- ings.
placements. Furthermore, maximum lateral displacements are
typically checked near the end of the design process for ser- SIMPLIFIED MODEL OF MULTISTORY BUILDING
viceability limits, by comparing the computed displacements
to an allowable upper limit on the maximum interstory drift. During the preliminary design stage of a building using dis-
Lateral displacements are typically computed as the displace- placement-based design criteria, approximate analyses are use-
ments computed with a linear elastic analysis of the structure ful for a rapid estimation of the lateral stiffness required to
when subjected to code-specified (reduced) lateral forces mul- avoid certain maximum tolerable interstory drifts. Similarly,
tiplied by a displacement amplification factor that is intended during the evaluation of existing buildings, approximate anal-
to account for the inelastic deformation expected in the struc- yses can be useful in providing a rapid estimation of the max-
ture during severe earthquake ground motions. This approach imum interstory drift that can occur during a particular ground
has been criticized for being inconsistent, for underestimating motion. To be appropriate for these situations, the approximate
displacement demands, and for often relying on startlingly dif- analyses must be simple and relatively quick to implement.
ferent relationships between elastic and inelastic displacements Furthermore, the analyses need to provide relatively good re-
(Uang 1991; Priestley 1995). sults. Simplifications are usually introduced in the model, in
Recently there has been a growing interest in displacement- the loading, in the analysis method, or in a combination of
based design procedures in which lateral displacement de- these elements.
mands are used instead of lateral force demands (Moehle Several investigators have used simplified models of build-
1992). During the preliminary design of new buildings, or for ings that take into account only shear-type deformations, in
a rapid seismic evaluation of existing buildings, there is a need models usually referenced in the literature as ‘‘shear build-
for estimating the maximum lateral displacements that can oc- ings’’ (or stick models). However, some studies (Chopra and
cur in the building during the design earthquake ground mo- Cruz 1986; Uang and Maarouf 1993) have shown that a struc-
tion. In particular, the estimation of the maximum roof dis- tural model that neglects overall flexural deformations may
placement and of the maximum interstory drift ratio (IDR) produce significant errors in estimating the earthquake re-
(defined as the ratio of the maximum interstory drift to the sponse of buildings. Other studies (i.e., Khan and Sbarounis
interstory height) is useful in identifying the required capaci- 1964) have shown that even for buildings whose primary lat-
ties, particularly the required lateral stiffness, in order to eral-resisting system consists of shear walls, the use of a pure
achieve the desired performance of the building. flexural model is also inappropriate.
The objectives of this paper are (1) to present an approxi- As shown in Fig. 1, lateral deformations in buildings are
mate method to estimate lateral displacements and maximum usually a combination of lateral shear-type deformations and
interstory drifts in multistory buildings subjected to earthquake lateral flexural-type deformations. Consideration of combined
flexural and shear deformations in frame buildings was studied
1
Res. Engr., Nat. Ctr. for Disaster Prevention, CENAPRED, Av. Delfı́n by Blume (1968), who introduced the dimensionless parameter
Madrigal No. 665, México, D.F. 04360, Mexico. ␳, defined as the ratio of the sum of the stiffness of all the
Note. Managing Editor: Sashi K. Kunnath. Discussion open until Sep- beams at the midheight story of the frame to the sum of the
tember 1, 1999. To extend the closing date one month, a written request stiffnesses of all the columns at the same story. This parameter
must be filed with the ASCE Manager of Journals. The manuscript for
this paper was submitted for review and possible publication on June 16,
is a measure of the relative beam-to-column stiffness and
1997. This paper is part of the Journal of Structural Engineering, Vol. hence controls the degree of participation of lateral flexural
125, No. 4, April, 1999. 䉷ASCE, ISSN 0733-9445/99/0004-0417 – 0425/ and shear deformations in a moment-resisting frame building.
$8.00 ⫹ $.50 per page. Paper No. 16012. The case when ␳ equals zero represents a flexural-type
JOURNAL OF STRUCTURAL ENGINEERING / APRIL 1999 / 417

J. Struct. Eng. 1999.125:417-425.


FIG. 1. Overall Lateral Deformations in Buildings
FIG. 2. (a) Simplified Model of Multistory Building; (b) Inter-
acting Forces in Model
building deforming as shown in Fig. 1(a), in which the beams
do not impose restraint to joint rotations. When ␳ equals ⬁ the
Downloaded from ascelibrary.org by University of Leeds on 09/17/13. Copyright ASCE. For personal use only; all rights reserved.

structure becomes a shear building that deforms as shown in


Fig. 1(b), in which the joint rotations are completely restrained
and deformations occur only through double-curvature bend-
ing of columns. An intermediate value of ␳ represents a frame
that combines shear and flexural lateral deformations, as
shown in Fig. 1(c), where both beams and columns undergo
bending deformations with joint rotations. The influence of ␳
on the maximum base shear and the maximum overturning
moment at the base of buildings has been studied by Chopra
and Cruz (1986).
In the method presented here, the multistory building is
modeled as an equivalent continuum structure consisting of a
combination of a flexural cantilever beam and a shear canti-
lever beam deforming either in bending or shear configura-
tions, respectively. The flexural and shear cantilever beams are
assumed to be connected by axially rigid members that trans-
mit horizontal forces; thus, the flexural and shear cantilevers
in the combined system deflect laterally by the same amount
[Fig. 2(a)]. FIG. 3. Effect of Nondimensional Parameter a on Lateral Load
Distribution Computed with (5)
This type of model was first proposed by Khan and Sba-
rounis (1964) to evaluate the interaction of shear walls and
frames. In their study the estimation of lateral displacements where
and forces in the system was solved by using an iterative
method. Heidebrecht and Stafford Smith (1973) derived the ␣2 GA
= (4)
differential equation that controls the system and provided H 2 EI
closed-form solutions to the lateral displacement, bending mo-
ments, and shear forces under static triangular and uniform Similar to the parameter proposed by Blume, the dimen-
lateral load distributions. sionless parameter ␣ in (3) controls the degree of participation
If the system is subjected to a distributed load (i.e., force of overall flexural and shear deformations in the simplified
per unit length) w(z), then, as shown in Fig. 2(b), a distributed model of multistory buildings. A value of ␣ approximately
horizontal interaction force of magnitude q(z) and a concen- equal to zero represents a pure flexural model [Fig. 1(a)] and
trated force Q at the top are required to maintain the displace- a value equal to ⬁ corresponds to a pure shear model [Fig.
ment compatibility and force equilibrium between the shear 1(b)]. An intermediate value of ␣ corresponds to a multistory
and the flexural cantilevers. building that combines shear and flexural deformations [Fig.
The differential equation for shear in the flexural cantilever 1(c)].
element is given by (Heidebrecht and Stafford Smith 1973) In a building subjected to earthquake ground motions, the


H
distribution of lateral loads varies at every instant of time. It
d 3u is common to assume that the distribution of lateral forces
⫺EI = [w(z) ⫺ q(z)] dz ⫺ Q (1) along the height of the building is triangular. A more general
dz 3 z
lateral load distribution is proposed here as follows:
where u = lateral displacement; EI = flexural rigidity of the
flexural beam; and H = total height of the cantilever beam. 1 ⫺ e⫺a z/H
w(z) = Wmax (5)
The differential equation for shear in the shear cantilever beam 1 ⫺ e⫺a
is


H
where Wmax is the intensity of the distributed load at the roof;
du and a is a dimensionless parameter that controls the shape of
GA = q(z) dz ⫹ Q (2) the lateral load. As shown in Fig. 3, the extreme values of a,
dz z
zero, and ⬁ correspond to triangular and uniform load distri-
where GA = shear rigidity of the shear cantilever beam. butions, respectively. Similarly, a value of a equal to 2.13 cor-
Differentiating, summing, and dividing by EI (1) and (2) responds to an approximately parabolic lateral load distribu-
yields tion (Fig. 4).
The lateral displacement [i.e., the solution to (3)] when the
d 4u ␣2 d 2u w(z) building is subjected to a distributed lateral load (5) is given
4 ⫺ = (3)
dz H 2 dz 2 EI by
418 / JOURNAL OF STRUCTURAL ENGINEERING / APRIL 1999

J. Struct. Eng. 1999.125:417-425.


Downloaded from ascelibrary.org by University of Leeds on 09/17/13. Copyright ASCE. For personal use only; all rights reserved.

FIG. 4. Comparison of Parabolic Lateral Load Distribution to


That Computed with (5) When a Is Equal to 2.13 FIG. 5. Effect of Ratio of Shear to Flexural Deformations on


Lateral Displacement Profile When Subjected to Triangular
Wmax H 4 z z Load Distribution
u(z) = C1 sinh ␣ ⫹ C2 cosh ␣ ⫹ C3 e⫺a z/H
EI(1 ⫺ e⫺a) H H

冉冊 册
2
z z
⫹ C4 ⫹ C5 ⫹ C6
H H (6)

where C1 – C6 are constants that depend on the particular so-


lution of (3) and on the boundary conditions. For fixed at the
bottom and free at the top boundary conditions, these constants
are given by

␣2e⫺a ⫺ a2e⫺a ⫺ a3 ⫹ a␣2 ⫺ ␣2


C1 = (7)
a␣3(a2 ⫺ ␣2)

a2e⫺a ⫺ ␣2e⫺a ⫹ a3 ⫺ a␣2 ⫹ ␣2 sinh ␣


C2 =
a␣3(a2 ⫺ ␣2) cosh ␣
␣2e⫺a ⫹ a2 ⫺ ␣2 1

␣4(a2 ⫺ ␣2) cosh ␣ (8)

⫺1 FIG. 6. Effect of Lateral Load Distribution on Lateral Displace-


C3 = (9)
a2(a2 ⫺ ␣2) ment Profile

⫺1 APPROXIMATE EARTHQUAKE ANALYSIS


C4 = (10)
2␣2
In the response spectrum analysis of a multistory building
a2e⫺a ⫺ ␣2e⫺a ⫹ a3 ⫺ a␣2 the peak value of the nth-mode contribution of the lateral dis-
C5 = (11) placement of the jth floor, uj , and the interstory drift of the jth
a␣2(a2 ⫺ ␣2)
story, ⌬j , are computed (Chopra 1995) as
␣2e⫺a ⫺ a2e⫺a ⫺ a3 ⫹ a␣2 ⫺ ␣2 sinh ␣ ujn = u stjn Dn (13)
C6 =
a␣3(a2 ⫺ ␣2) cosh ␣
⌬jn = ⌬ Dnst
jn (14)
1 ␣2e⫺a ⫹ a2 ⫺ ␣2 1
⫹ 2 2 2 ⫺
where u and ⌬ are the modal static displacements and the
st
jn
st
jn
a (a ⫺ ␣ ) ␣4(a2 ⫺ ␣2) cosh ␣ (12) modal static interstory drifts in the jth floor and in the jth story,
Fig. 5 shows the lateral displacement normalized with re- respectively; and Dn is the displacement spectrum ordinate cor-
spect to the displacement at the roof for different values of the responding to the natural period Tn and damping ratio ␰ n . The
nondimensional parameter ␣ when subjected to a triangular modal static displacements and interstory drifts are given by
load distribution. For a given roof displacement, shear-type u stjn = ⌫n␾jn (15)
buildings (␣ = 30) can have displacements at midheight that
are nearly 100% larger than those in flexural-type buildings (␣ ⌬ = ⌫n (␾jn ⫺ ␾jn ⫺1)
st
jn (16)
⬇ 0). The effect of the lateral load distribution along the height where ␾jn = nth-mode shape value at the jth floor; and ⌫n is
of the building is shown in Fig. 6, where the lateral displace- the modal participation factor defined as
ment normalized with respect to the displacement at the roof


N
corresponding to a triangular and a uniform lateral load dis-
tribution are compared. It can be seen that the lateral loading mj ␾jn
Ln


⌫n =
j=1
pattern has a smaller effect than that of the degree of partici- = N (17)
Mn
pation of flexural and shear lateral deformations on the lateral mj ␾jn
2

displacement profile. j=1

JOURNAL OF STRUCTURAL ENGINEERING / APRIL 1999 / 419

J. Struct. Eng. 1999.125:417-425.


in which mj = lumped mass at the jth floor; and N = number
of floors in the building.
The modal static displacements and interstory drifts can be
computed using a static analysis of the building when sub-
jected to lateral forces sjn at the various floor levels given by
sjn = ⌫n mj ␾jn (18)
In the method presented here several approximations are
made: (1) only the contribution of the first mode is considered;
(2) the distribution of masses along the height of the building
is assumed to be uniform; (3) the lateral stiffness is assumed
to be uniform along the height of the building; and (4) in
computing the participation factor an approximate shape func-
tion is used instead of the first mode. Considering only the
Downloaded from ascelibrary.org by University of Leeds on 09/17/13. Copyright ASCE. For personal use only; all rights reserved.

contribution of the first mode is incorrect, however, several


pilot studies (Saiidi and Sozen 1981; Miranda 1991; Qi and
FIG. 7. Lateral Displacements Normalized by Spectral Dis-
Moehle 1991; Alonso et al. 1996; Collins et al. 1996) have placement in Building Subjected to Triangular Load Distribution
shown that for moderate levels of story ductility demands (i.e.,
smaller than five) this assumption leads to relatively good ap-
proximations of the response of many low and midrise build- flexural building (␣ ⬇ 0). Furthermore, for buildings that de-
ings. form laterally like a shear building (␣ = 30), a displacement
The approximate lateral displacement at the jth floor level equal to the spectral displacement occurs at approximately
is computed as z/H = 0.57, whereas for buildings where lateral flexural de-
formations dominate over lateral shear deformations (␣ ⬇ 0)
uj = ␤1␺j Sd (19) a lateral displacement equal to the spectral displacement is
expected to occur higher in the building (at approximately
where ␤1 is the approximate participation factor; ␺j is the as- z/H = 0.72).
sumed shape value at the jth floor; and Sd is the displacement For a continuum structure like that shown in Fig. 2(a), the
spectrum ordinate corresponding to the fundamental period of rotation (or slope) at any point along its height can be obtained
vibration of the building T1. The approximate participation fac- through its derivative, which can be used as an approximation
tor is given by to the interstory drift ratio IDR (interstory drift divided by the


N story height), that is
␺j uj⫹1 ⫺ uj du(z j)
⬇ ␪(z j) =


␤1 =
j=1
(20) IDRj = (23)
N hj dz
␺2
j
j=1 where z j = height of the jth floor; and hj = corresponding
interstory height.
where the assumed shape is given by The IDR at relative height z/H and normalized by the roof
u(z j) drift ratio (defined as the maximum roof displacement divided
␺j = ␺(z j) = (21) by the height of the structure) is given by
u(H )
du(z/H ) H
where z j = height of the jth floor (measured from the ground
level); N = number of stories in the building; H = total height dz u(H )
of the building; and u(z j) and u(H ) are the lateral displace- z z z
ments at the jth floor level and the roof computed with (3) C1 ␣ cosh ␣ ⫹ C2 ␣ sinh ␣ ⫺ C3 ae⫺a z/H ⫹ 2C4 ⫹ C5
H H H
using a predefined lateral load distribution pattern along the =
height of the building. For buildings with uniform story height, C1 sinh ␣ ⫹ C2 cosh ␣ ⫹ C3 e⫺a ⫹ C4 ⫹ C5 ⫹ C6
the total height, H, is equal to the number of stories, N, times (24)
the story height. where C1 – C6 are given by (7) – (12). Fig. 8 shows IDRs nor-
Substitution of (6) into (21) gives malized by roof drift ratios in buildings subjected to a trian-

冉冊 gular load distribution. It can be seen that for buildings with


2
z z z z
C1 sinh ␣ ⫹ C2 cosh ␣ ⫹ C3 e⫺a z /H ⫹ C4 ⫹ C5 ⫹ C6 ␣ = 30 the interstory drift demands near the bottom are ex-
H H H H
␺(z) =
pected to be 50% larger than the roof drift ratio, whereas near
C1 sinh ␣ ⫹ C2 cosh ␣ ⫹ C3 e⫺a ⫹ C4 ⫹ C5 ⫹ C6 the top of the building they are expected to be less than one-
(22) half the roof drift ratio. For buildings in which lateral flexural
deformations dominate over lateral shear deformations (␣ ⬇
where C1 – C6 are given by (7) – (12). Fig. 7 shows the maxi- 0), the largest interstory drift demands occur near the top of
mum lateral displacements normalized by the spectral dis- the building and are approximately 35% larger than the roof
placement when subjected to a triangular load distribution (a drift ratio. Buildings subjected to a triangular load distribution
= 0.01). It can be seen that for values of z/H smaller than 0.85 that combine shear and flexural lateral deformations with val-
the normalized lateral displacements increase as the overall ues of ␣ between 2 and 4 will experience IDRs approximately
shear deformations dominate over the lateral flexural defor- 30% larger than the roof drift ratio near midheight, and IDRs
mations (i.e., as ␣ increases), whereas for z/H larger than 0.85 similar to the roof drift ratio near the top of the building.
the opposite is true. For the same spectral displacement the
lateral displacement at midheight (z/H = 0.5) in a building ESTIMATION OF MAXIMUM LATERAL
where lateral shear deformations dominate over lateral flexural DEFORMATIONS
deformations (␣ = 30) is nearly twice the lateral displacement
at the same height in a building that deforms laterally like a During the preliminary design stage of a new structure or
420 / JOURNAL OF STRUCTURAL ENGINEERING / APRIL 1999

J. Struct. Eng. 1999.125:417-425.


Downloaded from ascelibrary.org by University of Leeds on 09/17/13. Copyright ASCE. For personal use only; all rights reserved.

FIG. 8. IDRs Normalized by Roof Drift Ratio in Buildings Subjected to Triangular and Uniform Load Distribution

for a rapid seismic evaluation of an existing building using a


performance-based design methodology, an estimation of the
maximum roof displacement and of the maximum interstory
drift ratio IDRmax may suffice to evaluate the required lateral
stiffness in a new building or to evaluate if the lateral stiffness
of an existing building is adequate in order not to exceed pre-
determined performance levels under different levels of earth-
quake ground motions.

Maximum Roof Displacement


Since the assumed shape ␺j corresponds to lateral displace-
ments normalized by the lateral displacement at the roof, (19)
reduces to

uroof = ␤1Sd (25)

From (25), it can be seen that the approximate participation FIG. 9. Effect of Number of Floors and of Dimensionless Fac-
factor ␤1 represents the ratio between the maximum roof dis- tor ␣ on ␤1
placement and the spectral displacement. Thus, this factor,
which is computed by substituting (22) into (20), can be con-
sidered as an amplification factor acting on the spectral dis-
placement in order to obtain an estimate of the maximum roof
displacement. Fig. 9 shows the value of ␤1 for buildings with
uniform story height as a function of the number of stories,
N, and of the dimensionless parameter, ␣, when subjected to
a triangular load distribution. For a single-story building the
factor ␤1 is equal to unity, which implies that, as expected, the
maximum roof displacement is equal to the spectral displace-
ment. It can be seen that, regardless of the value of ␣, the
amplification factor ␤1 increases with an increasing number of
stories; thus, the larger the number of stories the larger the
difference between the spectral displacement and the roof dis-
placement. Furthermore, for a given number of stories, the
difference between the spectral displacement and the roof dis-
placement increases as ␣ decreases (i.e., as flexural deforma-
tions increase with respect to shear deformations). FIG. 10. Effect of Lateral Loading Pattern on ␤1
The effect of the lateral loading pattern is shown in Fig. 10,
where the amplification factor ␤1 is given for a triangular load placement computed using spectral analysis (13), (15), and
pattern (a = 0.01) and a uniform load pattern (a = 2,000). It (17).
can be seen that the loading pattern has practically no effect Expressions similar to (25) have been proposed in the past.
on buildings where shear deformations are negligible with re- Moehle (1984) used the first mode as the assumed displace-
spect to flexural deformations (␣ ⬇ 0) and has a small effect ment shape ␺j to estimate roof displacements in reinforced
on ␤1 for values of ␣ smaller than 30; thus, if the value ␣ is concrete buildings. Assuming a uniform story height and uni-
correctly selected, the estimation of the maximum roof dis- form mass distribution and a triangular (i.e., linear) displace-
placement computed with (25) yields a very good approxi- ment shape, Algan (1982) proposed an amplification factor ␤1
mation of the contribution of the first mode to the roof dis- given by
JOURNAL OF STRUCTURAL ENGINEERING / APRIL 1999 / 421

J. Struct. Eng. 1999.125:417-425.


Downloaded from ascelibrary.org by University of Leeds on 09/17/13. Copyright ASCE. For personal use only; all rights reserved.

FIG. 11. Comparison of Ratio of Roof Displacement to Spectral Displacement, ␤1, Computed with (20) and (22) to That Computed
with (26)

冘冉 冊
N
j numerically. Solutions to (29) for triangular (a = 0.01), uni-
N form (a = 2,000), and nearly parabolic (a = 2.13) lateral load
j=1 3N

冘冉 冊
␤1 = = (26) patterns are shown in Fig. 12. As shown in this figure, for
N
j
2
2N ⫹ 1 buildings where the overall flexural deformations dominate
j=1 N over shear deformations (␣ ⬇ 0), the maximum IDR is ex-
pected to occur near the top of the building, and for buildings
where N is the number of stories in the building. where overall shear deformations dominate over flexural de-
Fig. 11 shows a comparison between the amplification fac- formation (␣ > 20) the maximum IDR is expected to occur at
tor ␤1 computed using (20) and (22) with a = 0.01 (triangular a height smaller than one-fifth of the total height of the build-
load distribution) with that computed using (26). It can be seen ing. Furthermore, it can be seen that the effect of the lateral
that (26) leads to an overestimation of the roof displacement loading pattern increases as the value of ␣ increases. As shown
for buildings where shear deformations dominate over flexural in this figure, the lateral loading pattern has a relatively small
deformations (␣ > 4) and to a slight underestimation of the effect on the height at which the maximum IDR occurs.
roof displacement for buildings where flexural deformations The amplification factor ␤2 can be computed by substituting
dominate over shear deformations (␣ ⬇ 0). It can be seen that the solution of (29) into (28). Fig. 13 shows the amplification
a linear displacement shape, as it is done in many building factor ␤2 as a function of the dimensionless parameter ␣ when
codes, can provide relatively good estimates of the maximum subjected to triangular (a = 0.01), uniform (a = 2,000), and
roof displacement, particularly for buildings where shear de- nearly parabolic (a = 2.13) lateral loading patterns. For build-
formations dominate over flexural deformations; however, a ings where shear deformations dominate over flexural defor-
linear displacement shape assumes a uniform IDR and thus mations (␣ > 20), the amplification factor ␤2 is significantly
provides a poor estimation of the maximum IDR. larger than for buildings that deflect like a flexural cantilever
beam (␣ ⬇ 0), which means that shear-type buildings have
Maximum IDR larger concentrations of interstory drifts than flexural-type
An estimate of IDRmax can be computed as buildings. However, there is not a monotonic increase of the
amplification factor ␤2 as ␣ increases, but there is an optimum
uroof u(H ) value of ␣ that minimizes the ratio of the maximum IDR and
IDRmax = ␤2 = ␤2 (27)
H H the roof drift ratio. The optimum values of ␣ are 2.18, 1.87,
where ␤2 is given by

␤2 = max 冋 du(z) H
dz u(H )
册 (28)

From (27) it can be seen that the factor ␤2 is the ratio be-
tween the maximum IDR and the roof drift ratio. Thus, this
factor can be considered as an amplification factor acting on
roof drift ratio in order to obtain the maximum IDR. This
factor takes into account that usually the distribution of in-
terstory drifts along the height of the building is not uniform
and accounts for the concentrations of interstory drifts.
The height at which the maximum IDR will occur can be
computed from
d 2u(z/H ) z z
= C1␣2 sinh ␣ ⫹ C2 ␣2 cosh ␣ ⫹ C3 a2e⫺a z /H ⫹ 2C4 = 0
dz 2 H H
(29)
FIG. 12. Effect of Lateral Loading Pattern on Height at Which
where C1 – C6 are given by (7) – (12). Eq. (29) can be solved Maximum IDR Occurs

422 / JOURNAL OF STRUCTURAL ENGINEERING / APRIL 1999

J. Struct. Eng. 1999.125:417-425.


Downloaded from ascelibrary.org by University of Leeds on 09/17/13. Copyright ASCE. For personal use only; all rights reserved.

FIG. 14. Inelastic Displacement Ratios for Firm Alluvium Sites


FIG. 13. Effect of Lateral Loading Pattern on ␤2
where T = period of vibration of the structure; and ␮ is the
and 2.02 for triangular, uniform, and nearly parabolic load displacement ductility ratio. The ratio of inelastic to elastic
distributions, respectively. displacements computed with (31) is also shown in Fig. 14,
For values of ␣ smaller than five the lateral loading pattern where it can be seen that this equation leads to very good
produces changes smaller than 10% in the amplification factor approximations of the inelastic displacement ratio.
␤2 . However, this difference can increase to 44% for a value Eq. (31) can also be used for structures on rock; however,
of ␣ equal to 20 and a uniform load distribution. Thus, the it should not be used for structures built on very soft soil
effect of the lateral load pattern on the concentration of in- deposits. The reader is referred to Alonso et al. (1996) for
terstory drifts is more pronounced for shear-type buildings inelastic displacement ratios that can be used for buildings on
than for flexural-type buildings. very soft soils.
The use of (31) has been derived from single-degree-of-
Lateral Deformation Demands in Inelastic Systems freedom systems and thus it only gives an estimate of the
global response of multiple-degree-of-freedom systems. Fur-
Eqs. (25) and (27) provide good approximations of the max- thermore, Miranda (1991) has shown that the displacement
imum roof displacements and the maximum IDR in a building profile ␺j remains constant while the structure remains elastic.
responding linearly; thus, they can be directly used in the per- However, as inelastic deformations increase the displacement
formance-based design of buildings for those performance lev- profile changes and an increase in concentration of interstory
els in which the structure is expected to remain elastic or drifts occurs, which will produce an increase in the amplifi-
nearly elastic. However, for structures that are allowed to be- cation factor ␤2 . The possible increase in ␤2 depends primarily
have inelastically, as in the case of buildings of normal im- on three factors: (1) the number of stories, (2) the maximum
portance, during severe earthquake ground motions, a further story displacement ductility ratio, and (3) the type of mecha-
modification is necessary to estimate the maximum deforma- nism. In general, the larger the number of stories and/or the
tion demands. larger the maximum story displacement ductility ratio, the
To explore to what extent inelastic displacement demands larger the increase in the amplification factor ␤2 . In buildings
can be predicted using linear elastic analyses, Miranda (1991, whose mechanism involves column hinging, the amplification
1993) computed the ratio of the maximum inelastic to maxi- will be larger than that in buildings designed according to a
mum elastic displacement in 31,000 single-degree-of-freedom strong-column weak-girder philosophy.
systems undergoing five levels of inelastic deformation, when For the preliminary design of buildings to be constructed
subjected to 124 earthquake ground motions recorded on dif- according to a strong-column weak-girder philosophy, (27) can
ferent soil conditions. In this study, it was concluded that this be modified empirically as follows:
ratio depends on the period of vibration of the system, on the
level of inelastic deformation, and on the local soil conditions. uroof
Results of this study for firm alluvium sites are shown in Fig. IDRmax = ␤2 ␤4 (32)
H
14. It can be seen that for a displacement ductility ratio of 2
the maximum inelastic displacements are approximately equal where ␤4 is given by
to the maximum elastic displacements for periods greater than
0.4 sec, whereas for a displacement ductility ratio of 6 such ␮ N
␤4 = 1 ⫹ ⫹ (33)
an assumption is only valid for periods greater than approxi- 30 200
mately 1.1 sec. To account for this increase in displacement
where ␮ = maximum story displacement ductility ratio; and N
demands for short-period buildings with inelastic behavior, it
= number of stories. For existing buildings, or for the design
is proposed that the maximum inelastic roof displacement be
of a new building in which a preliminary design exists, a better
computed as
estimation of ␤4 can be obtained through nonlinear static anal-
uroof = ␤1␤3 Sd (30) yses (Miranda 1991, 1996; Collins et al. 1996).

where ␤3 is the inelastic displacement ratio defined as the ratio EVALUATION OF PROPOSED METHOD
of the maximum inelastic displacement, ui , to the maximum
elastic displacement, ue , which for firm alluvium sites can be To evaluate the effectiveness of the simplified method pre-
estimated as sented here to estimate maximum drift demands in buildings

冋 冉 冊 册
⫺1 subjected to earthquakes, it was used with the 10-story steel
ui 1 moment-resisting frame (MRF) building shown in Fig. 15,
␤3 = = 1⫹ ⫺1 exp(⫺12T␮⫺0.8 ) (31)
ue ␮ which was designed according to the 1994 Uniform Building
JOURNAL OF STRUCTURAL ENGINEERING / APRIL 1999 / 423

J. Struct. Eng. 1999.125:417-425.


Downloaded from ascelibrary.org by University of Leeds on 09/17/13. Copyright ASCE. For personal use only; all rights reserved.

FIG. 16. Comparison of Roof Displacements Computed Using


Detailed Two-Dimensional Nonlinear Time-History Analyses
with Those Computed with Simplified Method

multistory buildings responding primarily in the fundamental


mode when subjected to earthquakes was presented. The pro-
cedure is based on a simplified model of multistory buildings
that consists of a combination of a flexural cantilever beam
FIG. 15. Ten-Story Steel MRF Building Used to Test Proposed
Method
and a shear cantilever beam. The simplified model was used
to investigate the ratio of the spectral displacement to the roof
Code (Uniform Building Code 1994). The structure is assumed displacement and of the ratio of the maximum IDR to the roof
to be located in zone 4 with soil type 2 and has a uniform drift ratio.
lateral stiffness along its height. A 5% damping ratio is as- The effect of the number of stories, of the lateral loading
sumed. The weight associated with each floor in the frame is pattern, and of the ratio of overall flexural and shear defor-
1,107 kN and the fundamental period is 1.9 sec. mations on the ratio of spectral displacement and maximum
Since in this case the structure is known, the factor ␣ was roof displacement was studied. It was concluded that the dif-
computed as the one that minimizes the difference in the nor- ference between the spectral displacement and the maximum
malized lateral displacement profile resulting from a linear roof displacement increases with the number of stories. Fur-
elastic analysis of the building applying an arbitrarily chosen thermore, for a given number of stories, the difference between
intensity of lateral loads with a triangular load distribution spectral displacement and maximum roof displacement in-
with that computed using (22) with a factor a = 0.01 (trian- creases as overall flexural deformations increase with respect
gular load distribution). For this building the resulting value to shear deformations. The lateral loading pattern has a very
of ␣ is 16. small effect on the ratio of spectral displacement to the max-
Maximum roof displacements and maximum IDRs were imum roof displacement.
computed using nonlinear two-dimensional step-by-step time For buildings where overall shear deformations dominate
history analyses when subjected to the following ground mo- over flexural deformations, the ratio of the maximum IDR to
tions: (1) N – S component of the May 18, 1940, El Centro, the roof drift ratio is significantly larger than for buildings
Calif. earthquake; (2) N21E component of the Taft July 21, where overall shear deformations are negligible with respect
1952, Kern County, Calif. earthquake; and (3) N – S compo- to overall flexural deformations, which means that MRF build-
nent of the Corralitos October 17, 1989, Loma Prieta, Calif. ings are likely to have larger concentrations of interstory drifts
earthquake. The building was analyzed iteratively by scaling than flexible-type buildings (shear wall buildings). The effect
the ground motion intensity until the maximum story displace- of the lateral loading pattern on the concentration of interstory
ment ductility ratio in the building was equal to one, two, drifts is more pronounced for shear-type buildings than for
three, and four. Maximum roof displacements and maximum flexural-type buildings.
IDRs were then computed using the proposed simplified The method was tested by comparing deformation demands
method. computed using detailed nonlinear time-history analyses with
Fig. 16 shows a comparison of the maximum roof displace-
ment estimated with the simplified method (30) with that com-
puted with nonlinear step-by-step time history analyses. It can
be seen that the simplified method produces very good esti-
mates of the maximum roof displacements for displacement
ductilities of one and two; however, the method tends to over-
estimate the maximum roof displacement for story displace-
ment ductilities of three and four. Fig. 17 shows a comparison
of the maximum IDR estimated with the simplified method
(32) to that computed with step-by-step time history analysis.
It can be seen that despite the fact that only the contribution
of the first mode is considered and despite significant inelastic
deformations, the simplified method proposed here produced
very good estimates (with errors smaller than 26%) of the
maximum IDR.
SUMMARY AND CONCLUSIONS FIG. 17. Comparison of Maximum IDRs Computed Using De-
tailed Two-Dimensional Nonlinear Time-History Analyses with
A procedure to estimate roof displacements and IDRs in Those Computed with Simplified Method

424 / JOURNAL OF STRUCTURAL ENGINEERING / APRIL 1999

J. Struct. Eng. 1999.125:417-425.


those resulting from the proposed method. The method yields Heidebrecht, A. C., and Stafford Smith, B. (1973). ‘‘Approximate analysis
approximate deformation demands that can be considered as of tall wall-frame structures.’’ J. Struct. Div., ASCE, 99(2), 199 – 221.
Khan, F. R., and Sbarounis, J. A. (1964). ‘‘Interaction of shear walls and
accurate enough for the preliminary design of new buildings
frames.’’ J. Struct. Div., ASCE, 90(3), 285 – 335.
and for a rapid evaluation of existing buildings. Miranda, E. (1991). ‘‘Seismic evaluation and upgrading of existing build-
ings,’’ PhD thesis, University of California at Berkeley, Calif.
ACKNOWLEDGMENTS Miranda, E. (1993). ‘‘Evaluation of site-dependent inelastic seismic de-
This work has been partially supported by the Institute of Engineering sign spectra.’’ J. Struct. Engrg., ASCE, 119(5), 1319 – 1338.
of the National Autonomous University of Mexico. The assistance of Miranda, E. (1996). ‘‘Assessment of the seismic vulnerability of existing
graduate student Perla Santa-Ana in carrying out the analyses to evaluate buildings.’’ Proc., 11th World Conf. on Earthquake Engrg., Paper 513,
the proposed procedure is greatly appreciated. Prof. Kevin R. Collins Elsevier, Oxford, England.
from the University of Michigan and two anonymous reviewers provided Moehle, J. P. (1984). ‘‘Strong motion drift estimates for R/C structures.’’
constructive comments to improve this paper. J. Struct. Engrg., ASCE, 110(9), 1988 – 2001.
Moehle, J. P. (1992). ‘‘Displacement-based design of RC structures sub-
APPENDIX. REFERENCES jected to earthquakes.’’ Earthquake Spectra, 8(3), 403 – 428.
Qi, X., and Moehle, J. P. (1991). ‘‘Displacement design approach for
Algan, B. B. (1982). ‘‘Drift and damage considerations in earthquake-
reinforced concrete structures subjected to earthquakes.’’ Rep. No.
Downloaded from ascelibrary.org by University of Leeds on 09/17/13. Copyright ASCE. For personal use only; all rights reserved.

resistant design of reinforced concrete buildings,’’ PhD thesis, Univer-


sity of Illinois, Urbana, Ill. EERC/UCB-91/02, Earthquake Engineering Research Center, Univer-
Alonso, J., Miranda, E., and Santa-Ana, P. (1996). ‘‘Inelastic displace- sity of California, Berkeley, Calif.
ment demands for structures built on soft soils.’’ Proc., 11th World Saiidi, M., and Sozen, M. A. (1981). ‘‘Simple nonlinear seismic analysis
Conf. on Earthquake Engrg., Paper 40, Elsevier, Oxford, England. of R/C structures.’’ J. Struct. Div., ASCE, 107(5), 937 – 951.
Blume, J. A. (1968). ‘‘Dynamic characteristics of multi-story buildings.’’ Priestley, M. J. N. (1995). ‘‘Myths and fallacies in earthquake
J. Struct. Div., ASCE, 94, 337 – 402. engineering — Conflicts between design and reality.’’ ACI Spec. Publ.,
Chopra, A. K., and Cruz, E. F. (1986). ‘‘Evaluation of building code 157, 231 – 254.
formulas for earthquake forces.’’ J. Struct. Engrg., ASCE, 112, 1881 – Uang, C.-M. (1991). ‘‘Establishing R (or Rw) and Cd factors for building
1899. seismic provisions.’’ J. Struct. Engrg., ASCE, 117, 19 – 28.
Chopra, A. K. (1995). Dynamic of structures, theory and applications to Uang, C.-M., and Maarouf, A. (1993). ‘‘Safety and economy considera-
earthquake engineering. Prentice-Hall, Englewood Cliffs, N.J. tions of UBC seismic force reduction factors.’’ Proc., 1993 Nat. Conf.,
Collins, K. R., Wen, Y. K., and Foutch, D. A. (1996). ‘‘Dual-level seismic Central United States Earthquake Consortium, Memphis, 121 – 130.
design: A reliability-based.’’ Earthquake Engrg. and Struct. Dyn., 25, Uniform Building Code. (1994). International Conference of Building Of-
1433 – 1467. ficials, Whittier, Calif.

JOURNAL OF STRUCTURAL ENGINEERING / APRIL 1999 / 425

J. Struct. Eng. 1999.125:417-425.

You might also like