Sie sind auf Seite 1von 12

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/268801636

Load-Deformation Responses of Slender Reinforced Concrete Walls

Article  in  Aci Structural Journal · January 2004


DOI: 10.1016/j.engstruct.2017.02.050

CITATIONS READS

71 1,126

2 authors:

Leonardo M. Massone John W. Wallace


University of Chile University of California, Los Angeles
35 PUBLICATIONS   428 CITATIONS    151 PUBLICATIONS   2,038 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Utilizing Remote Sensing to Assess the Implication of Tall Building Performance on the Resilience of Urban Centers View project

Experimental evaluation of the performance of concrete coupling beams subjected to inelastic demands under wind loading View project

All content following this page was uploaded by Leonardo M. Massone on 15 November 2017.

The user has requested enhancement of the downloaded file.


ACI STRUCTURAL JOURNAL TECHNICAL PAPER
Title no. 101-S12

Load-Deformation Responses of Slender Reinforced


Concrete Walls
by Leonardo M. Massone and John W. Wallace

A detailed review of experimental data obtained from select slender tests to assess the relative contributions of flexural and shear
reinforced concrete (RC) wall tests was conducted to assess the deformations to inelastic lateral displacements. An important
relative contributions of flexural and shear deformations to feature of the study is to assess the accuracy and consistency
inelastic lateral displacements. An important feature of the study is of the experimental results, including any coupling between
to assess the accuracy and consistency of the experimental results, inelastic flexural and shear deformations, as well as to provide
including any coupling between inelastic flexural and shear defor-
mations, as well as to provide vital data to support the development
vital data to support the development and calibration of nonlinear
and calibration of nonlinear models. Based on these studies, it was models. The evaluation of appropriate values of lateral load
found that commonly used approaches, which rely on diagonal stiffness for use in common analysis methods is also reviewed.
displacement transducers mounted within the yielding region of
the wall, tend to overestimate shear distortions by as much as 30%. EXPERIMENTAL PROGRAM
An approach to correct the results based on the use of vertical Experimental results were obtained for six, approximately
displacement transducers within the yielding region is evaluated
1/4-scale wall specimens. The walls tested included three
and found to produce consistent results for the tests evaluated. The
use of 4 to 6 pairs of vertical displacement transducer pairs was
walls with a rectangular cross section (one with an opening),
found to be effective. Evaluation of the test results also indicates two walls with a T-shaped cross section, and one wall with a
coupling between inelastic flexural and shear deformations, barbell-shaped cross section with an opening. An overview
despite nominal shear strengths of approximately twice the shear of these studies is provided in the following paragraphs, with
force applied during the test. more detailed information concerning the walls without
openings (Thomsen and Wallace 1995, 2004) and the
Keywords: deformation; shearwall; stiffness; test; wall. walls with openings (Taylor, Cote, and Wallace 1998)
presented elsewhere.
INTRODUCTION
Reinforced concrete (RC) structural walls are commonly Test specimen information
used to resist the actions imposed on buildings due to earth- The walls were 3.66 m (12 ft) tall and 102 mm (4 in.) thick,
quake ground motions because they provide substantial with web and flange lengths of 1.22 m (4 ft). Floor slabs
strength and stiffness. The strength is required to limit damage were provided at 0.914 m (3 ft) intervals over the height of
in more frequent earthquakes and sufficient wall deformation the T-shaped walls. Typical material properties were selected for
capacity or ductility must be provided to ensure the lateral design, that is, fc′ = 27.4 MPa (4 ksi) and fy = 414 MPa (60 ksi).
load capacity is maintained during the inelastic response Boundary vertical steel consisted of 8 No. 3 (Ab = 71 mm2
expected during stronger, less frequent, earthquakes. The [0.11 in. 2]) bars, whereas web bars were deformed No. 2
large lateral-load stiffness commonly associated with structural (Ab = 32 mm2 [0.049 in.2]). Detailing requirements at the
walls limits the deformations of the lateral force resisting boundaries of the wall specimens were evaluated using the
system as well as the deformations imposed on the other displacement-based design approach presented by
structural and nonstructural elements. The use of structural Wallace (1994, 1995). Special boundary elements were
walls is also a popular and effective means to rehabilitate provided over the bottom 1.22 m (4 ft) of each wall. A capacity
deficient existing construction. design approach was used to avoid shear failure (Table 1),
Where structural walls are used, their behavior dominates and favorable anchorage conditions existed for the vertical
the lateral load strength and stiffness of the structure; therefore, reinforcement anchored within the pedestal at the base of the
it is essential that the strength, stiffness, and deformation wall. Reinforcing details for rectangular wall specimens
capacities of the walls be well understood. Relatively few, RW1 and RW2 are shown in Fig. 1 and the T-shaped wall
well-documented, experimental studies exist in the literature cross sections TW1 and TW2 are in Fig. 2.
to assess these attributes. Of particular importance is the
need to provide a comprehensive assessment of the load versus
Materials
deformation response for flexure and shear. The availability
of this information would be valuable in evaluating or calibrating Design compressive strengths were 27.6 MPa (4000 psi);
existing models for nonlinear response of structural walls, as however, strengths at the time of testing ranged from 28.7 to
well as providing vital data for assigning appropriate stiffness 58.4 MPa (4150 to 8460 psi), with mean compressive strengths
values for linear and nonlinear modeling.
ACI Structural Journal, V. 101, No. 1, January-February 2004.
MS No. 02-453 received December 26, 2002, and reviewed under Institute publication
OBJECTIVES policies. Copyright © 2004, American Concrete Institute. All rights reserved, including
The objectives of this study were to conduct a detailed review the making of copies unless permission is obtained from the copyright proprietors.
Pertinent discussion including author’s closure, if any, will be published in the November-
of experimental data obtained from select slender RC wall December 2004 ACI Structural Journal if the discussion is received by July 1, 2004.

ACI Structural Journal/January-February 2004 103


horizontally to a reaction wall 3.81 m (12.5 ft) above the base
Leonardo M. Massone is a PhD student in the Department of Civil Engineering at
UCLA, Los Angeles, Calif. He received his BS from the University of Chile in 1999
of the wall. Out-of-plane support was provided to prevent
and his MS in civil engineering from UCLA in 2003. His research interests include twisting of the wall specimen during testing. The displacement
analytical and experimental studies of reinforced concrete elements and systems. histories applied to Specimens RW2 and TW2 are shown in
John W. Wallace, FACI, is an associate professor of civil engineering at UCLA. He is Fig. 5. The displacement history for RW1 was similar to
a member of ACI Committee 318-H, Seismic Provisions; 335, Composite and Hybrid RW2, except that the four additional cycles—two at 1% and
Structures; 369, Seismic Repair and Rehabilitation; 374, Performance-Based Seismic
Design of Concrete Buildings; E 803, Faculty Network Coordinating Committee; and
Joint ACI-ASCE Committee 352, Joints and Connections in Monolithic Concrete Table 1—Test summary
Structures. His research interests include response and design of buildings and
bridges to earthquake actions, and laboratory and field-testing of structural FMAX
components and systems. ≡ VMAX / FMAX δ at
Specimen Reinforcement FN* VN VMAX† VN /FN FMAX δ/hw‡

at the base of the wall specimens (0 to 0.91 m [0 to 3 ft]) of ID Boundary Web kN kN kN (6)/(5) (6)/(4) mm (9)/hw
31.6, 34.0, 43.6, and 41.7 MPa (4580, 4925, 6330, and 6050 psi) (1) (2) (3) (4) (5) (6) (7) (8) (9) (10)
for Specimens RW1, RW2, TW1, and TW2, respectively. Peak 6.35
130.7 275.8 141.4 0.51 1.08 69.9 0.0191
stress was reached for a cylinder compressive strain of RW1 8 to 9.5§ at –130.7 275.8 –148.6 0.54 1.14 –58.9 –0.0161
||
approximately 0.002 for all specimens (Fig. 3(a)). More 190.5
detailed material information is available elsewhere (Thomsen 6.35
130.7 275.8 158.3 0.57 1.21 79.2 0.0217
RW2 8 to 9.5 at
and Wallace 1995). Figure 3(b) plots the measured stress-strain 190.5 –130.7 275.8 –157.5 0.57 1.21 –76.2 –0.0208
relations for the reinforcement used in the studies. 6.35
TW1 8 to 9.5 at 193.2 275.8 195.3 0.71 1.01 58.7 0.0160
–355.8 275.8 –290.9 1.05 0.82 –45.5 –0.0124
Testing and instrumentation 190.5
The wall specimens were tested in an upright position 6.35
at 178.7 342.5 189.7 0.55 1.06 88.9 0.0243
(Fig. 4). An axial load of approximately 0.10Ag fc′ was applied TW2 8 to 9.5
140.0 –342.6 342.5 –363.0 1.06 1.06 –82.3 –0.0225
at the top of the wall by hydraulic jacks mounted on top of *Lateral force required to reach MN at wall base.
the load transfer assembly. The axial stress was applied †Maximum applied lateral force (positive/negative).
prior to imposing lateral displacements, and it held constant ‡Top of wall drift (h = 3658 mm = 12 ft).
w
throughout the duration of each test. Cyclic lateral displacements §No. of vertical bars—bar diameter in mm.

were applied to the walls by a hydraulic actuator mounted ||Bar diameter in mm at bar spacing in mm.

Fig. 1—Specimens RW1 and RW2—boundary reinforcing details.

Fig. 2—Reinforcing details—Specimens TW1 and TW2.

104 ACI Structural Journal/January-February 2004


(a)

Fig. 4—Specimen test setup—RW1 and RW2.

(b)
Fig. 3—Material stress-strain relations: (a) concrete; and
(b) reinforcement.

two at 1.5% drift—were not applied after applying the first


two cycles at these drift levels. All measurements were read
by a computer data acquisition system at 32 points throughout
each cycle (Fig. 5). The applied displacement histories for TW1
and TW2 are identical to RW2 (up until failure was reached).
Instrumentation was used to measure displacements, loads, Fig. 5—Applied displacement history.
and strains at critical locations for each wall specimen. Four
wire potentiometers (WPs) were mounted to a rigid steel
reference frame to measure lateral displacements at 0.91 m
(3 ft) intervals over the wall height. A linear potentiometer was
also mounted horizontally on the pedestal to measure any
horizontal slip of the pedestal along the strong floor. Two
additional linear potentiometers were mounted vertically at
each end of the pedestal to measure rotation caused by uplift of
the pedestal from the strong floor. Shear deformations were
measured through the use of WPs mounted on the bottom two
stories (in an X configuration) of each specimen (Fig. 6).
Axial (vertical) displacements at the wall boundaries were
measured using two WPs mounted directly to the wall ends
and were used to calculate wall base rotations (by dividing
the difference in relative axial displacements by the distance
between the potentiometers). Linear variable differential
transformers (LVDTs) were mounted vertically (Fig. 6) over a
gage length of 229 mm at various locations along the Fig. 6—Instrumentation provided.
depth of each wall so that axial strains and curvatures
could be determined. The rectangular walls were instrumented EXPERIMENTAL RESULTS
with LVDTs spaced along the length of the wall, whereas the General observations about behavior and failure modes
T-shaped walls had 5 LVDTs mounted along the web and four are provided by Thomsen and Wallace (1995, 2004). Also,
mounted along the outside face of the flange. The strains in the an assessment of the displacement-based design approach
reinforcing steel were measured through the use of strain gages. used to determine the required transverse reinforcement at
All three types of reinforcement were monitored (longitudinal, the wall boundaries is presented by Thomsen and Wallace
uniformly distributed web, and transverse boundary steel) (2004). Lateral force versus deformation responses for flexure
at various locations (refer to Thomsen and Wallace 1995). and shear are studied in this paper.

ACI Structural Journal/January-February 2004 105


with an extreme fiber strain concrete compressive strain of
0.003 and an arbitrary displacement using the material relations
plotted in Fig. 3. A section analysis assuming a plane section
was used to find the moment for first yield of reinforcement.
The associated yield displacement was computed by integrating
the curvature over the wall height using the calculated
moment-curvature relation. The nominal moment capacity
was computed for an extreme fiber strain concrete compressive
strain of 0.003. The bilinear predictions are plotted in Fig. 7
along with the experimental determined relative displacement
between the top of the wall and the wall-pedestal interface
for Specimens RW2 and TW2. The plots reveal that lateral
load capacities of the test walls are slightly greater than nominal
values. Maximum experimentally applied lateral loads are
(a) listed in Table 1 and range from 1.01 to 1.21 times the
nominal values, except for Specimen TW1 for the flange in
tension, which only reached a value of 0.82 due to intentionally
poor detailing provided at the web boundary (Detail A,
Fig. 2(a)), which led to premature buckling of the vertical
web and boundary reinforcement.
The analytical comparisons shown in Fig. 7 are based on
including only flexural deformations, that is, deformations
due to shear and anchorage/bond are excluded. The wall
aspect ratio (3) and the wall shear strength were selected to
ensure that elastic and inelastic deformations would be
dominated by flexure. Also, the wall vertical reinforcement
was continuous over the wall height and favorable conditions
were provided for the anchorage of vertical reinforcement
(web and boundary) extending into the base pedestal. An
(b) examination of the test results revealed that top displacement
due to anchorage slip was negligible. The significance of shear
Fig. 7—Lateral load versus top displacement relation:
deformations on wall top displacement is examined further.
(a) Specimen RW2; and (b) Specimen TW2.
Load-displacement response: shear deformations—Given
Wall nominal moment and shear capacities that the nominal shear capacity was not exceeded (refer to
The nominal moment capacities were computed using Table 1), the displacement at the top of the wall due to elastic
stress-strain relations developed for 27.6 MPa (4 ksi) shear deformations was estimated using the design material
compressive strength concrete and for reinforcement with a properties as
nominal yield of 414 MPa (60 ksi) (Fig. 3). Confined concrete
was incorporated using the model proposed by Saatcioglu
δ shear =  ---------- ( h w )
F
and Razvi (1992) as detailed by Thomsen and Wallace (1)
 GA ′
(2004). The lateral loads required to reach the nominal
moment capacity for each specimen are given in Table 1.
Two values are given for the T-shaped walls because the where F is the applied lateral force; G is the shear modulus;
moment capacities vary depending on the direction of the and A′ is the shear area. The top displacement due to elastic
applied load. The nominal shear capacities were computed for shear deformations is relatively small. For example, for RW2
nominal materials using Eq. (21-10) of ACI 318-02 and are also with F = 150 kN (33.7 kips), G = 13,100 MPa (1,900 ksi), A′ =
provided in Table 1. The nominal shear capacity is sufficient to (5/6) × (Aweb) = 0.103 m2 (160 in.2), and h = 3.81 m (150 in.),
allow the wall to reach the nominal moment capacity in all cases the top displacement due to shear deformations is 0.42 mm
except TW1, in which premature failure was anticipated due to (0.017 in.) or only 0.01% of the wall height. This estimate
poor boundary detailing (Thomsen and Wallace 1995). would approximately double if cracking is considered;
however, in either case, the contribution of elastic shear defor-
Load-displacement response mations to top displacement has little impact on the
The load versus top displacement responses of the test analytical relations.
specimens were computed using analytical relations. Shear and flexural deformations: evaluation of test data—
Comparisons between measured results and analytical results Instrumentation was provided on the test specimens to allow
are made for deformations due to flexure and shear to assess the determination of flexural and shear deformations over
the ability of the analytical relations to predict the experi- the first two stories of the specimens (Fig. 6). Detailed
mental results, as well as to assess the relative contributions evaluation of the measured responses is presented and
of flexural and shear deformations. discussed in the following subsections.
Load-displacement response: flexural deformations—A
bilinear approximation to the lateral load-top displacement Models for flexural and shear deformations
relation for each wall was computed using two points: the A common way to evaluate the average shear deformation
moment and displacement associated with first yield of the for structural wall tests is to use an X configuration of displace-
boundary vertical reinforcement and the moment associated ment gages (for example, Thomsen and Wallace [1995];

106 ACI Structural Journal/January-February 2004


Shiu et al. [1981]; Oesterle et al. [1976, 1979]; and Goodsir
[1985]). For the wall tests evaluated, WPs were placed diag-
onally over the first story from 0 to 0.91 m (0 to 3 ft) and the
second story from 0.91 to 1.83 m (3 to 6 ft), as well as vertically
along the wall boundary from 0 to 0.91 m (0 to 3 ft) (Fig. 6).
Based on the configuration of displacement gages, average
shear deformations are determined with the model of Fig. 8.
In Fig. 8(a), the lateral displacement U is the same for each
side unless higher order terms are considered. If the vertical
displacements V1 and V2 are such that the element center of
rotation is located in the element center, which occurs for spe-
cial cases (for example, uniformly distributed curvature), the
diagonal lengths are invariant and the average shear de-
Fig. 8—X configuration for symmetric elements. Flexural
formations measured using an X configuration are not affected
and shear behavior.
by flexural deformations (Fig. 8(b)).
For RC walls as well as other RC elements, the center of
rotation is not always located in the element center; therefore,
the average shear deformations computed with the X config-
uration must be corrected. Some authors have already accounted
for this effect by correcting their measurements (for example,
Farrar and Baker 1990, 1993; Hiraishi 1984; Salonikios 2002).
For example, Salonikios (2002) assumed that the flexural
lateral displacement is equal to the difference of vertical
displacement, which implies that the center of rotation is at
the bottom of the element. Farrar and Baker (1990 and
1993) used a different approach where the flexural displacement
at the top of a one-story wall specimen was computed assuming
that the centroid of the inelastic curvature distribution did not
change from the centroid for the elastic curvature distribution.
Hiraishi (1984) included a factor α, where α corresponds to the Fig. 9—Flexural model accounting vertical displacement.
fraction of the story height to the center of equivalent flexural
rotation measured from the base of the element or story, where Us is the average shear displacement; h is the story
to correct measured values ( α =1.0 is consistent with the height; D1meas and D2meas are the diagonal lengths for the
assumption used by Salonikios). deformed X configuration; V1 and V2 are the vertical
For this study, the geometry shown in Fig. 9 is used to displacements at the top of the story level; and Uf is the
compute the average shear deformations. The undeformed flexural lateral displacement.
geometry of the original rectangular element is outlined by First story total lateral displacement—The lateral displace-
broken lines. The element deformed by shear is shown as a ment over the first story (0.0 to 0.91 m) was measured using WPs
shaded rhomboid, and the combined shear and flexural mounted horizontally between a rigid reference frame mounted
deformations are shown in solid lines. According to the on the strong floor and the test specimen (Fig. 6). In addition,
geometry shown in Fig. 9, the diagonal length associated potentiometers were mounted on the foundation pedestal to
with pure shear displacement (D1shear and D2shear) can be monitor pedestal rotation and lateral slip so that the influence of
used to estimate the total average shear displacement Us as these deformations could be removed from the total first story
lateral displacement to determine the relative first story lateral
displacement. The lateral displacement over the first story can
shear
2
2 shear
2
2 also be computed from the diagonal and vertical measurements
D1 – h – D2 –h obtained during testing by rearranging Eq. (3) as
Us = ----------------------------------------------------------------------
- (2a)
2
Utotal ≡ Us + Uf = (4)
= (2b)
2 2
meas 2 meas 2
D1 – ( h + V2 ) – D2 – ( h + V1 )
– ( h + V 2 ) – U f2 –  D 2 – ( h + V 1 ) + Uf 1
2 2
meas 2 meas 2 ------------------------------------------------------------------------------------------------------
-
D1
  2
------------------------------------------------------------------------------------------------------------------------------------------
2
The results obtained using the two independent approaches
Noting that Uf1 = Uf2 = Uf, that is, the lateral displacement for determining the lateral displacement over the first story
due to flexure is equal on each element side, then are compared to assess the consistency of the measurements
as well as to assess the relative contributions of shear and
flexural deformations. Results are presented for three walls:
meas
2
2 meas
2
2 RW2, TW2, and SRCW1. Specimen SRCW1 is a 4.88 m
D1 – ( h + V2 ) – D 2 – (h + V1 ) (16 ft) tall, 1.22 m (4 ft) long, 152 mm (6 in.) thick RC wall
Us = ------------------------------------------------------------------------------------------------------
- – Uf (3)
2 with both structural steel W-sections and deformed reinforce-

ACI Structural Journal/January-February 2004 107


over the second story is only about 1/4 that measured in the first
story. The corrected relation for the second level results in a
slightly worse correlation then using the uncorrected relation,
apparently due to the smaller displacement values (diagonal and
vertical) and scatter associated with the measured data.
The results presented in Fig. 10 indicate that the experi-
mental results obtained are reliable, as the first story lateral
displacement results obtained using the two independent
approaches are in close agreement. The results also indicate
that the approach commonly used to estimate first story shear
deformations overestimates the lateral displacement by 16 to
31%, whereas the proposed model results in estimates that
are between 3% low to 11% high. Flexural deformations
typically accounted for about 70% of the first story lateral
displacement. To obtain the aforementioned comparisons,
the center of rotation for the flexural deformations was assumed
to be 1/3 of the distance from the base to the first story level
(that is, α = 2/3). The importance of this assumption is
addressed in the discussions that follow.
First story lateral displacement—flexural deformations—The
analysis of measured data from the diagonal WPs and reinforce-
ment strain gages indicates that yielding was primarily limited to
the first story; however, the distribution of the curvature over the
first story is unknown because only a single gage was used on
each side of the wall to derive the average rotation over the first
Fig. 10—Comparison of total displacement: (a) RW2; (b) TW2; story. That is, the first story rotation was obtained as
(c) Wall SRCW1; (c) first floor; and (d) second floor.
V1 – V2
θ = -----------------
- (5)
ment used as boundary longitudinal reinforcement (Cherlin l
2003). Given the wall aspect ratio (hw /lw) = 4 and the provided
nominal wall shear strength (Table 1), wall behavior is where l is the horizontal distance between the gages. To
dominated by flexural deformations. Provided instrumentation obtain the contribution of the flexural deformations (curvature)
is similar to that used for the other walls evaluated, except to the first story and top level lateral displacements, the location
vertical WPs were provided over both the first story (0 to 1.22 m of the centroid of the curvature distribution (center of
[0 to 4 ft]) and the second story (1.22 to 2.44 m [4 to 8 ft]). rotation) over the first story must be assumed. The flexural
Results obtained for the three specimens are plotted in Fig. 10. displacement for the first story is represented by
The relation plotted in Fig. 10 as uncorrected refers to calculating
the first story lateral displacement by summing the flexural
(vertical wire pots) and shear (diagonal wire pots) deformations Uf = αθ h 1 (6)
without correcting the shear deformations to account for the
influence of the vertical displacements on the diagonal where θ is the first story rotation obtained from Eq. (5); h1 is the
measurements. The relation plotted as corrected makes height of the first story; and α is the distance from the top of the
this correction using the model of Fig. 9. A 45-degree first story to the centroid of the curvature distribution. Thus, α
line on Fig. 10 indicates perfect agreement between the varies from 0.5 for the rectangular distribution to 0.67 for
two independent approaches for obtaining the first story the triangular distribution. In the results presented earlier, a
lateral displacement. A third relation plots only the contribution value of 0.67 was assumed based on the use of this value by
of the flexural deformations to the first story lateral displacement others (Thomsen 1995; Ali 1990). The influence of the value
by assuming the center of rotation is located 1/3 of the distance selected for α is addressed in the following paragraphs.
from the base (that is, α = 2/3). The relationship between elastic bending moment, bending
Results for RW2, plotted on Fig 10(a), reveal several stiffness, and curvature for either uncracked or cracked
important findings. First, the displacement obtained with the linear RC sections is represented by
uncorrected relation exceeds the displacement obtained with
the external instrumentation by about 20%. This is consistent 2
M- = -----------
d Uf-
with expectations, as the shear deformations are expected to φ = ----- (7)
EI 2
be overestimated with the uncorrected measurements. Excellent dx
agreement between the two approaches is obtained by correcting
the shear deformations. Flexural deformations account for where φ is the section curvature; M is the bending moment;
roughly 70% of the first story lateral displacement; however, it EI is the flexural stiffness; and Uf is the flexural displacement.
is important to note that they account for a much larger The flexural stiffness EI varies during the elastic stage of the
percentage of the top displacement. Results presented for wall testing from the initial uncracked section value to the
TW2 (Fig. 10(b)) and the first story of SRCW1 (Fig. 10(c)) fully cracked section value. Equation (7) can be extended
show similar trends. Results presented for the second story into the inelastic range if an equivalent (secant) stiffness
of SRCW1 (Fig. 10(d)) indicate that lateral displacement value is used. For the wall tests being reviewed, the moment

108 ACI Structural Journal/January-February 2004


Fig. 11—Gross, cracked, α = 0.755; 0.67; 0.585; 0.5, Fig. 12—Shear stiffness, different approaches: (a) RW2;
normalized flexural secant stiffness: (a) RW2; and (b) TW2. and (b) TW2.

over the wall height includes the contribution of: 1) the


applied load times the height to the base of the wall; 2) the
eccentricity of the applied axial load; and 3) the second order
effects due to the applied axial load acting through the
applied lateral displacement. In general, the curvature
distribution over the first story can be represented as


φ = ( 3 α – 1 ) ⋅ ------ (8)
h1

From Eq. (8), it is apparent that the value selected for α


influences the secant stiffness. Envelope values of normalized
secant stiffness (EIsec /EIg) for four different values of α are Fig. 13—RW2 displacements at 1st and 2nd floor: (a) flexural;
plotted in Fig. 11 for Walls RW2 and TW2. The figures show and (b) shear.
a relatively small variation of secant stiffness for α-values
between 0.6 and 0.7, which is a reasonable range for α for the 2 2
inelastic range. Given these results, use of an α-value of 0.67 D1
meas 2
– h – D2
meas
–h
2

in the earlier sections of this paper is appropriate. Figure 11 = --------------------------------------------------------------------


-+
2
also provides a useful reference for effective stiffness values as
a percent of EIg to use for nonlinear analysis approaches
such as the capacity spectrum approach. 1 V1 – V2 V1 – V2
--- ⋅ -----------------
- ⋅ h – α ⋅ -----------------
-⋅h
Recalling the original expression for shear displacement 2 l l
from Eq. (2), the relations for the original and the corrected
X configuration shear displacements are Thus

Us_Xcorrected = Us_Xoriginal +  1
--- – α ⋅ θ ⋅ h (12)
2 2
meas 2 meas 2
D1 – h – D2 –h
Us_Xoriginal = --------------------------------------------------------------------
- (9) 2 
2
Equation (12), reported by Hiraishi (1984), reveals that for
Us_Xcorrected = (10) the case for α = 0.5, the expressions for the original and
corrected X configuration produce the same result. For values of
α > 0.5, which is common, the uncorrected (original) shear
2 2
meas 2 meas 2 displacement is overestimated.
D1 – ( h + V2 ) – D2 – ( h + V1 )
------------------------------------------------------------------------------------------------------
- – Uf Shear force versus deformation relations–effective stiffness—
2 The elastic shear stiffness, A′G, can be determined using the
measured shear displacement from Eq. (1) as
Terms V1 and V2 are small relative to h; therefore, a Taylor’s
series expansion of Eq. (10) including only first order F⋅h
terms produces A ′ G = ------------1- (13)
δs

Us_Xcorrected = (11) The shear displacement, including corrections, is derived


as the difference between the total lateral displacement and
flexural displacement for α = 0.67. Relations for various
– h – --- ⋅ V 2 –  D 2 – h – --- ⋅ V 1
2 2
meas 2 h meas 2 h
D1 drift ratios, normalized to the elastic shear stiffness, are
l  l  shown in Fig. 12. The results presented indicate a signifi-
-----------------------------------------------------------------------------------------------------------------
- – Uf
2 cant reduction in shear stiffness from the elastic shear stiffness,

ACI Structural Journal/January-February 2004 109


Table 2—RW2—SRW1: Shear and flexural
displacements at different story levels
RW2 1st floor 2nd floor
U shear U shear U shear
Xcorrected, Xoriginal, U flex, Xoriginal, U flex,
P, kN mm mm mm/mm mm mm/mm
0.00 0.00 0.00 0.00 0.00 0.00
43.3 — — 0.23 — 0.57
72.3 — 0.12 0.65 — 0.97
70.6 — 0.12 0.66 — 0.99
104.8 0.21 0.55 1.33 0.08 2.33
126.6 0.44 0.92 1.88 0.94 2.69
123.6 0.44 0.92 1.88 0.93 2.64
135.3 0.73 1.36 2.45 — 2.35
132.3 0.83 1.59 2.95 — 2.65
142.4 1.88 3.06 4.66 0.96 3.43
141.0 2.26 3.68 5.61 1.06 4.27
139.3 2.17 3.59 5.60 0.98 3.54
142.8 2.71 4.11 5.55 0.98 4.30
152.7 3.38 5.19 7.21 1.05 3.98
150.5 3.45 5.47 8.04 1.30 4.90
158.2 4.41 6.76 9.39 1.36 6.33
141.4 4.38 6.78 9.56 1.48 6.04

SRW1 1st floor 2nd floor 3rd floor Fig. 14—SRCW1: (a) flexural displacements at 1st, 2nd,
U U U and 3rd floor; shear displacements at 1st, 2nd, 3rd floor:
shear U shear U shear U shear U (b) Xoriginal; and (c) Xcorrected.
Xcor- U shear flex, Xcor- U shear flex, Xcor- Xorigi- flex,
rected, Xorigi- mm/ rected, Xorigi- mm/ rected, nal, mm/
P, kN mm nal, mm mm mm nal, mm mm mm mm mm occurring over the first story). Summary results for uncorrected
0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 and corrected shear deformations and normalized flexural
64.1 — 0.02 0.51 — 0.00 0.12 — 0.00 0.30 deformations are presented in Table 2.
1.49 0.10 0.81 1.40 0.33 0.52 0.60 –0.01 0.10 0.73 Plots of story shear versus flexural deformation for RW2
157.0 0.28 1.06 1.56 0.42 0.61 0.62 0.03 0.15 0.72 (Fig. 13(a)) reveal that: 1) the cracked stiffness obtained
211.3 0.79 1.94 2.29 1.13 1.48 1.12 0.48 0.41 0.98 from a moment versus curvature analysis approximates the
244.7 0.85 2.33 2.94 1.27 1.68 1.32 1.10 0.39 1.12
effective stiffness prior to yield very well; 2) yielding occurred at
291.6 1.61 3.98 4.70 1.74 2.28 1.75 1.32 0.66 1.43
a lateral force close to that associated with the lateral load to
reach the nominal moment; and 3) yielding was limited to
306.4 1.96 5.40 6.82 1.72 2.35 2.04 1.54 0.94 1.86
primarily the first story. For shear (Fig. 13(b)), the plots reveal
307.2 3.33 7.68 8.64 2.21 3.04 2.68 1.54 0.97 2.04
that: 1) inelastic shear behavior occurred in the first story despite
290.4 4.32 8.53 8.39 2.51 3.44 2.98 1.54 0.98 2.04 a nominal shear capacity of approximately twice the applied
303.7 6.87 11.90 10.07 3.34 4.59 3.99 1.63 1.08 2.12 story shear; 2) inelastic deformations were limited to essentially
282.8 5.23 10.44 10.49 3.43 4.53 3.55 1.57 1.00 1.98 the first story; and 3) the elastic shear stiffness represents the
Notes: — corresponds to unavailable points. measured shear stiffness in regions where flexural yielding
Flexural displacements are normalized to first yield expected deformation value
at wall bottom. was not observed (that is, the second level). The results for
shear behavior clearly demonstrate the coupling of inelastic
which is consistent with the relatively large contribution of shear and flexure, that is, inelastic flexural deformations
shear (~30%) to the first story lateral displacement. appear to have led to inelastic shear deformations.
Shear and flexural force versus deformation relations— Results for Specimen SRCW1 are presented in Fig. 14.
Further insight into the behavior of the wall specimens is Additional vertical boundary and diagonal WPs were provided
obtained by investigating the lateral force versus the deformation on Specimen SRCW1; therefore, it is possible to directly
behavior for shear and flexural force. This information may determine the shear force versus deformation relations for
also be valuable for calibrating nonlinear models used for story levels one, two, and three. Plots of story shear versus
shear walls, such as the multiple-vertical-line-element flexural deformation (Fig. 14(a)) reveal findings similar to those
(MVLE) model (refer to Orakcal, Conte, and Wallace 2002). for RW2, with the exception that slip between the structural steel
Figure 13 and 14 plot the story shear versus the flexural and section and the concrete appears to have contributed significantly
shear deformations within the first and second stories for to a loss of stiffness in the first story level. For shear (Fig. 14(b)
Specimen RW2 and the first to third stories for Specimen and (c)), the plots reveal that: 1) inconsistent results are obtained
SRCW1, respectively. For the second story level of RW2, the for elastic shear stiffness unless the results are corrected; and 2)
value for shear displacement was obtained using the X pattern inelastic shear behavior occurred in the first story despite a
without corrections as vertical measurements were not available nominal shear capacity of approximately twice the applied story
to make the correction, and the flexural deformations were taken shear. The consistency of the corrected shear stiffness values for
as the total displacement minus the displacement due to the first to third story levels provides high confidence that
shear and rotation (pedestal rotation and flexural rotation the instrumentation worked effectively and the method used to

110 ACI Structural Journal/January-February 2004


correct the measurements to remove the influence of flexural
deformations from the shear behavior is appropriate. Given this
confidence, the observed inelastic shear deformations in the first
story level, and also to a lesser degree in the second story level,
clearly demonstrate the coupling on inelastic flexural and shear
deformations. Also, the flexural and shear relations show almost
the same yielding point. Although this observation appears in
prior works (Oesterle et al. 1979; Takayanagi, Derecho, and
Corley 1979), the failure to correct the shear force deformation
relation makes the prior findings slightly less reliable.
Figure 15 plots the cyclic story shear versus shear deformation
relation for the first story of Specimen RW2 and SRCW1.
These relations provide valuable data for the calibration Fig. 15—Cycles of shear displacements at 1st floor: (a) RW2;
of shear force versus deformation relations for macro- and (b) SRCW1.
models, such as the MVLE model (for example, Orakcal,
Conte, and Wallace 2002). Early MVLE models incorporated
origin-oriented hysteresis relations for shear (Vulcano and
Bertero 1987), which do not appear to be appropriate based
on the results presented in Fig. 15.
Shear force versus deformation relations—improved
instrumentation schemes—The findings presented would be
strengthened if the α-value could be estimated directly from the
experimental results. This would require additional instru-
mentation; specifically, the use of additional vertical sensors
at each boundary of the wall specimen that would allow
the determination of the distribution of curvature over the
wall height to be determined (Fig. 16). Given the costs associated
with using additional instrumentation, the number of transducers Fig. 16—Instrumentation for shear and flexural measurement
required to sufficiently determine the α-value is of interest. at one story.
Therefore, a parameter study was conducted where V2k and V1k
represent the vertical displacement of a pair of transducers over average curvature was assumed to be constant over each
gage length xk such that sensor length xk and the magnitude of the curvature distribution
was assumed to decrease as a step function with height h
above the base of the wall, that is
∑ k = 1 ( V2 – V1 ) ⁄ l
k k
θ =
N
(14)
k+1
φ pattern k+1 k 1
----------------- = C ⇒ φ pattern = C ⋅ φ pattern (18)
k
φ pattern

∑k = 1 x
N k
h1 = (15)
where C represents the step increment between adjacent
transducers. Values for C were taken as 1.1, 1.5, and 5 to
represent a range of possible situations. Convergence to the
k k k k
φ = ( V2 – V1 ) ⁄ ( x ⋅ l ) (16) exact solution is achieved if a very large number of transducers
are used. The criteria used in this study to determine the
In general, equal transducer lengths would be used; therefore, xi number of transducers required to achieve exact solution is
= x j for all i and j, resulting in based on an error tolerance of 0.1% in the α-value when the
number of transducers is increased from using N*/2 to N* for a
given value of C. Accordingly, the exact α-value for a given
∑k = 1φ ⋅  k – 1--- value of C is
N k
 2
α = ------------------------------------ (17)
φ ⋅  k – ---
Σ N
* k 1
∑k = 1  2
k
N⋅ φ
N N*
k=1 C - C+1
α exact = ------------------------------------------- = ----------------- – -----------------------------------
-(19)
N ⋅Σ
N* *
C – 1 2 ⋅ N ⋅ (C – 1)
*
* N k
φ
k=1
For known values of α, θ, and h1, the correction for the
shear lateral displacement could be assessed directly from The relationship between the two adjacent curvature values
Eq. (12). The value α is influenced by two factors: the average for N transducers, for a given C, is
curvature over each sensor gage length φk and the number of
vertical sensors used N.
*
A parameter study is conducted to assess the importance N
q ⋅ ------ *
N
of these factors on the α-value. Hypothetical curvature distri- N
φ -
q+1 ------


q N
k N
butions are generated to assess the number of sensors required φ = φ ⋅ -----*- ⇒ -----------
q
= C (20)
to accurately determine the α-value. Several constraints were * N φ
N
k = ( q – 1 ) ⋅ ------ + 1
placed on the hypothetical curvature distributions. First, the N

ACI Structural Journal/January-February 2004 111


Fig. 17—α/α pattern using different number of transducer N under
ratio of curvature distribution C.
Therefore, the ratio between the α-value for an arbitrary transducers used for this experiment. Direct calculation of
number of transducers N, where N < N*, and the exact α-value, the α-value from the experimental derived curvature values
for a given value of C, is results (that is, computing the centroid of the curvature
distribution) in an α-value of αdata = 0.594. Unfortunately,
N
*
N
* additional test data are not available to assess the validity of
 ------N  N*  N
------  N* the approach presented; however, results obtained from the
2 ⋅ N ⋅  C – 1 ⋅ C –  C + 1 ⋅ ( C – 1 )
   
α - = --------------------------------------------------------------------------------------------------------------- (21)
two methods are in close agreement and suggest that using a
------------- moderate number of transducers (4 to 6) is sufficient to
α exact *
2 ⋅ N ⋅ (C – 1) ⋅ C – (C + 1) ⋅ (C – 1)
N* N*
obtain reliable results.

* SUMMARY AND CONCLUSIONS


N (C – 1)
× ------ ⋅ ----------------------
*
- Test results for RC and structural steel RC walls were
N N
 N ------  examined to establish force deformation relations and to
 C – 1 study the reliability of shear deformation measurements. In
  addition, a parametric study was conducted to determine
appropriate instrumentation schemes for yielding regions to
N* is arbitrarily set to 32 for this study; however, the selection allow accurate assessment of the contribution of flexural
of N* for the wall specimens tested will be addressed in deformations to story lateral displacement. Based on these
subsequent paragraphs. The results presented in Fig. 17 for studies, the following conclusions were reached:
N = 6 and 1 < C < 1.1; N = 8 and 1.1 < C < 1.5; and N = 12 1. Average shear distortion is commonly obtained by using
and 1.5 < C < 5.0 indicate that relatively few sensors are an X configuration of displacement transducers over the first
required to obtain a reasonable estimate of the α-value. For story level and, in some cases, for higher story levels. For
C < 1.5, the use of four vertical transducers produces reasonable slender wall tests using this configuration, the influence of the
results because the expected error relative to the exact solution differential vertical displacements at the edges of the wall
is generally less than 5%. should be accounted for in computing shear distortion, as
The results of the parametric study were used to assess test the measured distortions tend to overestimate the contribution
results obtained for an RC wall with embedded structural steel of shear deformations to wall lateral displacements. Results
boundary columns (Cherlin 2003). Specifically, test results are obtained without correction overestimated shear distortions at
evaluated for Specimen SRCW3, which had the same first story by 16 to 31%, whereas corrected results were between
longitudinal reinforcement and geometry as Specimen SRCW1 3% low and 11% high. Consistent results were obtained for the
described previously; however, the test was conducted with different wall tests by comparing first story level lateral
higher axial load (0.2Ag fc′ versus 0.1Ag fc′ for SRCW1) and displacements derived from two independent approaches;
slightly different boundary transverse reinforcement spacing 2. It is common practice to estimate the contribution of
was provided. Four displacement transducers were used over the flexural deformations to first story lateral displacement by
first story height of 1.22 m (4 ft), which is a reasonable assuming the center of rotation of the inelastic curvature
upper-bound estimate of the plastic hinge length (equal to distribution to be 1/3 of the distance from the base of the wall
the wall length). (α-value of 2/3). This assumption is common because sufficient
In the elastic range, the moment and curvature distributions instrumentation is not usually provided to determine the center
are known; therefore, selection of an α-value only applies of rotation directly using experimental measurements. An
where the yield curvature is exceeded. For SRCW3, the yield analytical study was conducted where key terms were varied
curvature was estimated from section analysis to be approx- to assess a reasonable distribution of vertical displacement
imately 0.0039/lw. Average curvature results were deter- transducers to determine α directly from experimental measure-
mined for the four pairs of transducers and adjacent values ments. The results indicated that four to six displacement trans-
were compared to assess the range of C-values from the test ducers at each wall boundary would be sufficient for the wall tests
results. An average C-value of 1.042 was determined; therefore, evaluated. Experimental results from one wall test where four
the results presented in Fig. 17 for 1.0 < C < 1.1 are appropriate. pairs of vertical displacement transducers were used supported
The value of N* was determined to be 32 based on the 0.1% this finding and suggested an α-value of 0.6 for this wall test;
criterion presented previously. 3. Secant stiffness values were evaluated from the wall
Based on C = 1.042 and N* = 32, the α-value and its ratio tests for flexure (EI) and shear (GA′). For a lateral drift ratio
to the exact solution are determined to be αexperiment = 0.599 at the top of the wall of 1 and 2%, secant flexural stiffness
and α/αexact = 0.989. Therefore, the estimated error for values of 0.096EIg and 0.047EIg and secant shear stiffness
the α-value is approximately 1% for the four pairs of vertical values of 0.110GA′ and 0.037GA′ were found for a wall

112 ACI Structural Journal/January-February 2004


specimen with a rectangular cross section (Specimen RW2). REFERENCES
These results may be useful for selecting secant stiffness ACI Committee 318, 1995, “Building Code Requirements for Structural
values for nonlinear static analysis procedures such as the Concrete (ACI 318-95) and Commentary (318R-95),” American Concrete
Institute, Farmington Hills, Mich., 369 pp.
capacity spectrum approach used in ATC-40; and
ACI Committee 318, 1999, “Building Code Requirements for Structural
4. Force versus deformation relations for flexure and shear Concrete (ACI 318-99) and Commentary (318R-99),” American Concrete
were derived for the first story levels from the test results, Institute, Farmington Hills, Mich., 391 pp.
where inelastic deformations were concentrated. Inelastic ACI Committee 318, 2002, “Building Code Requirements for Structural
shear and flexural deformations initiated at essentially the Concrete (ACI 318-02) and Commentary (318R-02),” American Concrete
same lateral displacement level, despite nominal shear Institute, Farmington Hills, Mich., 443 pp.
strengths of approximately twice the lateral force required to Ali, A., and Wight, J., 1990, “Reinforced Concrete Structural Wall with
Staggered Opening Configuration Under Reversed Cyclic Loading,” Report
initiate flexural yielding. Consistent results were obtained No. ECE-8603624, National Science Foundation, Arlington, Va., 241 pp.
for all three wall tests evaluated. The findings indicate a strong Cherlin, M., 2003, “Design of Steel Reinforced Concrete Shear Walls
coupling between inelastic flexural deformations and inelastic Subject to Seismic Loading,” master’s thesis, University of California, Los
shear deformations. The derived force versus deformation Angeles, Calif., 204 pp.
relations should aid in the calibration of wall macro-models, Farrar, C. R., and Baker, W. E., 1990, “Stiffness and Hysteretic Energy
Loss of a Reinforced-Concrete Shear Wall,” Experimental Mechanics,
such as the multiple-vertical-line element model. V. 30, No. 1, pp. 95-100.
Farrar, C. R., and Baker, W. E., 1993, “Experimental Assessment of
ACKNOWLEDGMENTS Low-Aspect-Ratio, Reinforced Concrete Shear Wall Stiffness,” Earthquake
The work presented in this paper was supported by funds from the Engineering and Structural Dynamics, V. 22, pp. 373-387.
National Science Foundation (NSF) under NSF Grant No. BCS-9112962, Goodsir, W. J., 1985, “Design of Coupled Frame-Wall Structures for
CMS-9632457, and CMS-9810012. This financial support from NSF is Seismic Actions,” Research Report 85-8, Department of Civil Engineering,
gratefully acknowledged. The test results evaluated were conducted by J. H. University of Canterbury, New Zealand, 359 pp.
Thomsen, now with SGH, Arlington, Mass., and B. L. Sayre, now with the
Hiraishi, H., 1984, “Evaluation of Shear and Flexural Deformations of
Los Angeles Department of Water and Power. The assistance of UCLA PhD
Flexural Type Shear Walls,” Bulletin of the New Zealand National Society
student K. Orakcal and B. L. Sayre are greatly appreciated. Opinions, findings,
for Earthquake Engineering, V. 17, No. 2, pp. 135-144.
conclusions, and recommendations in this paper are those of the authors and do
not necessarily represent those of the sponsor or others mentioned. Oesterle, R. G.; Fiorato, A. E.; Johal, L. S.; Carpenter, J. E.; Russell, H. G.,
and Corley, W. G., 1976, “Earthquake Resistant Structural Walls—Test of
Isolated Walls,” Report No. GI-43880/RA-760815, National Science
NOTATION Foundation, Arlington, Va., 315 pp.
A′ = shear area (taken as 5/6 of wall web area in all cases) Oesterle, R. G.; Aristizabal-Ochoa, J. D.; Fiorato, A. E.; Russell, H. G.;
Ab = boundary vertical steel area and Corley, W. G., 1979, “Earthquake Resistant Structural Walls—Test of
Ag = gross cross-sectional area Isolated Walls—Phase II,” Report No. ENV77-15333, National Science
C = constant curvature ratio between two consecutive trans- Foundation, Arlington, Va., 327 pp.
ducers based on pattern case
Orakcal, K.; Conte, J. P.; and Wallace, J. W., 2002, “Nonlinear Modeling
D1shear,
of RC Structural Walls,” Urban Earthquake Risk, Seventh U.S. National
D2shear = diagonal length associated with pure shear displacement
Conference on Earthquake Engineering (7NCEE), July 21-25, Boston,
EI = flexural stiffness
Mass., Paper No. 246, 10 pp.
EIg = gross section stiffness
EIsec = secant stiffness Saatcioglu, M., and Razvi, S. R., 1992, “Strength and Ductility of
F = applied lateral force Confined Concrete,” Journal of Structural Engineering, ASCE, V. 118,
fc′ = concrete compressive strength No. 6, pp. 1590-1607.
fy = steel yield stress Salonikios, T. N., 2002, “Shear Strength and Deformation Patterns of R/C
G = shear modulus Walls with Aspect Ratio 1.0 and 1.5 Designed to Eurocode 8 (EC8),”
h = element height Engineering Structures, V. 24, pp. 39-49.
h1 = height of first story Shiu, K. N.; Daniel, J. I.; Aristizabal-Ochoa, J. D.; Fiorato, A. E.; and
hw = wall height Corley, W. G., 1981, “Earthquake Resistant Structural Walls—Test of
l = element length (horizontal distance between gages) Walls with and without Openings,” Report No. ENV77-15333, National
M = bending moment Science Foundation, Arlington, Va., 120 pp.
N = number of transducers Takayanagi, T.; Derecho, A. T., and Corley, W. G., 1979, “Analysis of
N* = number of transducers for pattern case Inelastic Shear Deformation Effects in Reinforced Concrete Structural
U = element lateral displacement Wall Systems,” Nonlinear Design of Concrete Structures, CSCE-ASCE-
Uf = flexural lateral displacement ACI-CEB International Symposium, University of Waterloo, Ontario,
Us = average shear displacement Canada, pp. 545-579.
Us_ Taylor, C. P.; Cote, P. A.; and Wallace, J. W., 1998, “Design of Slender
Xcorrected = average shear displacement for corrected X configuration Reinforced Concrete Walls with Openings,” ACI Structural Journal, V. 95,
Us_ No. 4, Jul.-Aug., pp. 420-433.
Xoriginal = average shear displacement for original X configuration Thomsen, J. H., IV, and Wallace, J. W., 1995, “Displacement-Based
V1,V2 = vertical displacement measured on WPs at each boundary Design of RC Structural Walls: Experimental Studies of Walls with
of wall Rectangular and T-Shaped Cross Sections,” Report No. CU/CEE-95/06,
V2k, V1k = vertical displacement of segment at level k Department of Civil and Environmental Engineering, Clarkson University,
xk = height of vertical transducers at level k Potsdam, N.Y., 353 pp.
α = fraction of story height (from top) to center of equivalent
Thomsen, J. H., IV, and Wallace, J. W., 2004, “Displacement-Based
flexural rotation
Design of Slender RC Structural Walls—Experimental Verification,” Journal
αexact = factor α calculated for pattern case
of Structural Engineering, ASCE. (accepted for publication)
δinelastic = inelastic displacement at top of wall
δs = displacement of one story height due to elastic shear Vulcano, A., and Bertero, V. V., 1987, “Analytical Models for Predicting the
deformations Lateral Response of RC Shear Walls: Evaluation of their Reliability,” Report
δshear = displacement at top of wall due to elastic shear deformations No. UCB/EERC 87/19, Earthquake Engineering Research Center, 90 pp.
φ = section curvature Wallace, J. W., 1994, “A New Methodology for Seismic Design of RC
φk = average curvature over each sensor gage length at level k Shear Walls,” Journal of Structural Engineering, ASCE, V. 120, No. 3,
based on N transducers pp. 863-884.
φkpattern = average curvature at Level k for pattern case, based on N* Wallace, J. W., 1995, “Seismic Design of RC Shear Walls; Part I: New
transducers k: 1..N* Code Format,” Journal of Structural Engineering, ASCE, V. 121, No. 1,
θ = element story rotation Jan. 1995, pp. 75-87.

ACI Structural Journal/January-February 2004 113


View publication stats

Das könnte Ihnen auch gefallen