Вы находитесь на странице: 1из 10

ACI STRUCTURAL JOURNAL

TECHNICAL PAPER

Title no. 110-S10

Flexural Drift Capacity of Reinforced Concrete Wall with


Limited Confinement
by S. Takahashi, K. Yoshida, T. Ichinose, Y. Sanada, K. Matsumoto, H. Fukuyama, and H. Suwada
The flexural drift capacity of reinforced concrete (RC) walls is
discussed in this study based on the test results of 10 specimens.
The test parameters were wall length, thickness, detailing, and
axial force. The detailing of the ties did not satisfy the ACI 318-08
requirements. Each specimen had a column at one end where an
axial force was applied. All specimens failed in flexural compression after yielding of the longitudinal bars. The observed flexural
drift capacity was between 0.4 and 1.2%. A set of equations to
predict the drift capacity is proposed wherein the hinge zone length
is assumed to be 2.5 times that of the wall thickness.
Keywords: boundary element; compressive failure; confinement; drift
capacity; plastic hinge; reinforced concrete wall.

INTRODUCTION
ACI 318-081 requires special detailing for boundary
elements of reinforced concrete (RC) walls to prevent flexural compressive failure under seismic forces. One of the
approaches to the detailing is based on the displacementbased concept.2 If the compressive strain of concrete is
expected to be larger than 0.003, the compressive zone is
required to be reinforced according to the requirement of the
special boundary element for confinement. This requirement
is verified by Thomsen and Wallace,3 who tested walls with
rectangular- and T-shaped cross sections.
The Japanese Code4 prescribes the ductility of RC walls
based mainly on the ratio of the neutral axis depth to the wall
thickness. This prescription is based on several experimental
studies, including those by Tabata et al.,5 who tested RC walls
with rectangular cross sections and large shear-span ratios.
The plastic hinge length Lp is important for estimating drift
capacity. Researchers have proposed various approaches.
In the study by Wallace and Orakcal,2 which was the basis
of the seismic requirements of ACI 318-08,1 the plastic
hinge length was assumed as one half of the wall length
(Lp = lw/2). Tabata et al.5 assumed Lp = 0.3lw. On the other
hand, Kabeyasawa et al.6 idealized the wall, assuming that
the strain of each boundary element is uniform within
each story; this idealization is almost equivalent to the
assumption of Lp = h, where h is story height. Orakcal and
Wallace7 divided the wall into eight segments in the direction of the height; this idealization is almost equivalent to
the assumption of Lp = h/8. Paulay and Priestley8 assumed
that Lp = 0.20lw + 0.044a from the test results of cantilever walls, where a is shear span length. Takahashi et
al.9 showed that the prescription of the Japanese Code4 is
implicitly based on the assumption of Lp = 2.5t, where t is
wall thickness.
The objective of this paper is to propose a set of equations
to predict the drift capacity of RC walls based on the assumption of Lp = 2.5t. To verify this assumption, 10 specimens
were tested. The detailing of these specimens does not satisfy
the seismic requirements of ACI 318-08,1 but such detailing
ACI Structural Journal/January-February 2013

may be preferred for ease of construction. The test parameters were wall length, thickness, detailing, and axial force.
The details of this experiment are available elsewhere.10,11
RESEARCH SIGNIFICANCE
There are many RC buildings that do not satisfy the
requirements of ACI 318-08,1 including those in Chile. Most
of the wall damage caused by the 2010 Chile earthquake was
related to the configuration and reinforcement detailing of
wall boundary elements.12 The damage indicated that the
performance of these walls was brittle, as expected. On the
other hand, there may have been many buildings that resisted
the earthquake, although they did not satisfy the requirements of ACI 318-08.1
The ACI 318-081 requirements are quite strict about
the boundary element; it may be sufficient to ensure large
ductility of walls. However, the findings of this research may
lead to simpler detailing for walls where a relatively smaller
compressive strain is expected. The findings of this research
may also be used to evaluate the seismic capacity of buildings
with walls that do not satisfy the ACI 318-081 requirements.
EXPERIMENTAL PROGRAM
Specimens
Ten RC wall specimens with a boundary column on only
one side were prepared to investigate the differences in
deformation capacity. Figure 1 shows the cross sections of
the specimens used in this research. Each specimen is named
as follows:
1. Perpendicular end wall: The first letter of the specimens nameP or Nmeans with or without a perpendicular wall, respectively. For example, Specimen PM5 in
Fig. 1(g) has a perpendicular end wall 130 mm (5.1 in.) long
and 60 mm (2.4 in.) thick. All the perpendicular end walls
have single-layer reinforcement (D4 at 80 mm [3.15 in.],
where D4 is a deformed bar with a nominal diameter of 4 mm
[0.16 in.]) and do not have confinement.
2. The ratio of wall panel length to wall thickness (lwp/t):
The second letter of the specimens nameS, M, or
Lexpresses that the lwp/t ratio is 6, 12, or 18, respectively, where the wall panel length lwp is defined, excluding
the column. For example, the ratio of Specimen PM5 in
Fig. 1(g) is 1200/100 = 12.
3. The ratio of neutral axis depth to wall thickness (c/t):
The vertical arrows in Fig. 1 indicate the location of the
ACI Structural Journal, V. 110, No. 1, January-February 2013.
MS No. S-2011-062 received March 2, 2011, and reviewed under Institute
publication policies. Copyright 2013, American Concrete Institute. All rights
reserved, including the making of copies unless permission is obtained from the
copyright proprietors. Pertinent discussion including authors closure, if any, will be
published in the November-December 2013 ACI Structural Journal if the discussion
is received by July 1, 2013.

95

Susumu Takahashi is a PhD Student of architectural engineering at Nagoya Institute


of Technology, Nagoya, Japan.
Kazuya Yoshida is a Masters Student of architectural engineering at Nagoya Institute of Technology.
ACI member Toshikatsu Ichinose is a Professor of architectural engineering at
Nagoya Institute of Technology.
ACI member Yasushi Sanada is an Associate Professor of architecture and civil engineering at Toyohashi University of Technology, Toyohashi, Japan.
Kenki Matsumoto is a Masters Student of architectural engineering at Nagoya Institute of Technology.
ACI member Hiroshi Fukuyama is a Chief Research Engineer at the Building
Research Institute, Tsukuba, Japan.
Haruhiko Suwada is a Research Engineer at the Building Research Institute.

neutral axis, whose computation method will be shown in


a later section of the paper. The number of the specimens
name expresses the approximate ratio of c/t. For example,
the ratio of Specimen PM5 in Fig. 1(g) is 493/100 = 4.9.
Although the sections of Specimens NM5 and NM4 are
identical, as shown in Fig. 1(a), the locations of the neutral
axis are different because of the difference of axial forces, as
shown in a later section.
4. With or without crosstie in boundary element:
Figure 2 shows the detail of the boundary elements of the
specimens, except Specimens NM5, NM4, and NM2.
The horizontal bars have a 135-degree hook, as shown in
Fig. 2(a). The cap bars have a 90-degree hook at both ends,
as shown in Fig. 2(b), and the cap bars vertical spacing
is 35 mm (1.4 in.), as shown in Fig. 2(c). The crossties have
90- and 135-degree hooks, as shown in Fig. 2(b), and are
staggered with a spacing of 70 mm (2.8 in.), as shown in
Fig. 2(c). The crossties in Specimens NM5 and NM4 are
located at a spacing of 35 mm (1.4 in.), as shown in Fig. 3(a).

Specimen NM2 does not have crossties, as shown in


Fig. 3(b).
The detailing of the wall reinforcement for Specimen NM3
is shown in Fig. 4. The spacings of the horizontal and vertical
bars are 35 and 100 mm (1.4 and 4.0 in.), respectively. The
reinforcement details of the wall panels of the other specimens are the same as those in Fig. 4. Because the wall thickness of each specimen is different, the lateral and vertical
wall reinforcement ratios vary from 0.54% to 0.84% and
from 0.19% to 0.30%, respectively. The lengths of boundary
elements (220 mm [8.7 in.] in Fig. 2(a)) are longer than half
the length of the neutral axis in most specimens, as specified
by the seismic requirements of ACI 318-08.1 In the vertical
direction, crossties are provided from the bottom to one-third
of the clear height h, as shown in Fig. 4. This value of h/3
is much shorter than the requirement of ACI 318-08.1 The
spacing of the crossties (70 mm [2.8 in.] in most specimens)
does not satisfy the requirements of ACI 318-081 either. The
cross-sectional areas of the crossties vary from 15 to 31% of
ACI 318-08.1
The clear heights of Specimens NM4 and NM5 are
1000 mm (3.3 ft), whereas those of the other specimens are
1200 mm (4.0 ft), as shown in Fig. 4. Eight No. 3 (D10)
longitudinal bars are provided in the boundary element of
all specimens (Fig. 2(a)). Twelve No. 5 (D16) longitudinal
bars are provided in the boundary column, except that of
Specimen NL2, where eight No. 4 (D13) bars are provided
(Fig. 1(e)) so the longitudinal reinforcement ratio (2.5%) is
similar to that of the other specimens.
All specimens are designed to fail in flexure; the shearto-flexural-capacity ratios vary from 1.2 to 2.3, where
the flexural and shear capacities are calculated based on
the Architectural Institute of Japan (AIJ) standards13 and
ACI 318-08,1 respectively. The material properties of the
steel bars are indicated in Table 1, where fy is the yield
strength, fu is the tensile strength, and Es is the elastic
modulus. The material properties of concrete are indicated

Fig. 1Specimen sections. (Note: Dimensions in mm; 1 mm = 0.039 in.; No. 4 is D13;
No. 5 is D16.)
96

ACI Structural Journal/January-February 2013

Fig. 3Boundary element of Specimens NM5, NM4, and NM2.

Fig. 2Boundary element except for Specimens NM5, NM4,


and NM2. (Note: 1 mm = 0.039 in.; No. 3 is D10.)
in Table 2, where fc is the compressive strength, Ec is the
elastic modulus, and fr is the modulus of rupture.
Test setup
Figure 5 shows the test setup. Lateral force was applied by
a hydraulic jack to a stiff loading steel beam fastened to the
specimen. All specimens had stiff RC stubs at both the top
and bottom for fixing with the loading frame. No axial force
was applied for Specimen NM4. For the other specimens, two
vertical hydraulic jacks were force-controlled so the moment
around the center of the boundary column is zero, as shown
in Fig. 5. The amount of the axial force was approximately
20% of the axial capacity of the boundary column (fcAg),
where Ag is the gross cross-sectional area of the column.
The applied axial load was approximately 240 kN (54 kips)
for Specimen NL2, 400 kN (90 kips) for NM5, and 540 kN
(121 kips) for the other specimens. Horizontal load was
applied 2425 mm (8.0 ft) above the bottom of the wall panel
for NM5 and NM4 (Fig. 5). The height of the horizontal load
was 2525 mm (8.3 ft) for the other specimens. The shearspan ratio of NL2 is 2525/2000 = 1.26, which is the smallest.
The shear span ratio of NS3 is 2525/1020 = 2.48, which is
the largest.
OBSERVED DAMAGE
AND DEFLECTION COMPONENT
Figure 6 shows the lateral load-drift relationship of
Specimen NL2. Lateral drift R is defined as the ratio of
measured lateral displacement D to specimen height h. The
displacement was measured at the top of the clear height in all
specimens. During the positive loading (column in tension),
the maximum strength (530 kN [119 kips]) was observed at
a +1.2% drift level. The strength was 1.1 times the analytical
flexural strength. Between the drift levels of +2 and +3%,
strength decreased rapidly. During the negative loading
direction (column in compression), the maximum strength
ACI Structural Journal/January-February 2013

Fig. 4Elevation of Specimen NM3. (Note: Dimensions in


mm; 1 mm = 0.039 in.)
Table 1Material properties of reinforcing bars
Bar

fy, MPa

fu, MPa

Es, GPa

D4

411 (351)*

521 (544)*

173 (192)*

No. 3 (D10)

391 (376)*

469 (520)*

199 (188)*

No. 4 (D13)

367

503

183

No. 5 (D16)

389 (387)

559 (563)

180 (180)*

Numbers in parentheses indicate material properties for Specimens NM4 and NM5.
Notes: 1 MPa = 145 psi; 1 GPa = 145 ksi.

Table 2Material properties of concrete


Specimen
NS3
NM3
NM2
NM2

fc, MPa

Ec, GPa

fr, MPa

38.3

28.4

2.65

37.8

27.8

2.45

37.6

28.6

3.07

33.4

23.9

2.55

NL2
PL6
PM5
PM3
NM5
NM4

Notes: 1 MPa = 145 psi; 1 GPa = 145 ksi.

97

Fig. 5Loading setup.

Fig. 6Load versus drift of Specimen NL2.

Fig. 8Buckling of reinforcement.

Fig. 9Instrumentation of specimens.

Fig. 7Specimen NL2 at maximum drift.


(280 kN [63 kips]) was observed at a 1.5% drift level. The
strength was 1.2 times the analytical flexural strength.
Figure 7 shows Specimen NL2 at the end of the experiment at a 5% drift level. The spalling of concrete started at
the bottom right corner of the wall panel at a +2% drift level.
98

The spalled zone extended toward the column until a +3%


drift level. On the other hand, the concrete of the boundary
column slightly spalled during the negative loading but
not during the positive loading, even at a +5% drift level.
Figure 8 shows the buckling of the longitudinal bars (eight
No. 3 [D10]) in the boundary element at a 3% drift level. The
buckled bars fractured in tension between the drift levels of
2 and 3%.
Figure 9 shows the linear variable differential transducer
(LVDT) used to evaluate the flexural drifts.7 Figure 10 shows
ACI Structural Journal/January-February 2013

Fig. 11Load versus shear drift (deformation) of Specimen NL2.


(Note: 1 mm = 0.039 in.)

Fig. 10Load versus flexural drift of Specimen NL2.


the load-versus-flexural-drift relationship of Specimen NL2.
Flexural drift at 80% of maximum strength Vmax (the black
circle in Fig. 10) is defined as flexural drift capacity in this
paper. The hysteresis loops are more spindle-shaped than
those in Fig. 6. The strength degradation in Fig. 10 is milder
than that in Fig. 6, which will be discussed in the following.
Figure 11 shows the load-versus-shear-drift relationship
of Specimen NL2. Shear drift was obtained by subtracting
flexural drift from total drift. Shear drift increased after flexural yielding. The black triangles in Fig. 10 and 11 show
the load step at the onset of strength degradation. The shear
drift just before the strength degradation (1.25%) was larger
than the corresponding flexural drift (0.75%). Note that the
maximum applied shear force is much smaller than the shear
strength computed according to ACI 318-081 (the top broken
line in Fig. 11). The shear drift did not increase during the
degradation. This is the reason why the strength degradation
in Fig. 10 is milder than that in Fig. 6.
Figure 12 shows the horizontal slip along one of the
flexural cracks near the center of the wall when the shear
drift was 1.25% or the shear deformation was 15 mm
(0.59 in.) (the black triangle in Fig. 11). The observed slip
was 8.5 mm (0.33 in.), which was more than one half of
the total shear deformation (15 mm [0.59 in.]). The damage
and overall behavior of the other specimens were similar
to Specimen NL2, except that the slips along the flexural
cracks were smaller than those in NL2. To discuss the cause
of the slip, the compressive force of concrete C is defined in
Eq. (1).
C = N + Ast f y Asc f y

(1)

where N is applied axial load; Ast is the gross sectional area


of longitudinal bars in tension; Asc is the gross sectional
area of longitudinal bars in compression; and fy is the yield
strength of the longitudinal bar.
The variable Vmax in the horizontal axis of Fig. 13 indicates the maximum applied shear force. The vertical axis of
Fig. 13 shows the slip drift angle, which is defined as the sum
of the observed slips (Ss) just before the strength degradation (the black triangle in Fig. 11) divided by the clear height
h. Slip was larger in the specimens with larger Vmax/C ratios.
ACI Structural Journal/January-February 2013

Fig. 12Slip along flexural crack of Specimen NL2. (Note:


1 mm = 0.039 in.)

Fig. 13Lateral slip drift Ss/h versus Vmax/C.


Such a correlation is not obtained between the slip and the
average shear stress (Vmax/Ag, where Ag is the gross sectional
area of the specimen). Because the slip is not the focus of
this study, only flexural drift is discussed in the following.
The load-versus-flexural-drift relationships of Specimens
NM3 and PM3 are shown in Fig. 14(a) and 15(a) to inves99

Fig. 16Elastic and plastic deformations.

Fig. 14Load and strain versus flexural drift of Specimen NM3.

compressive ductility of this wall was very limited at the ultimate drift. Therefore, the contribution of the perpendicular
wall should be ignored in evaluating the drift capacity. The
vertical axes of Fig. 14(b) and 15(b) show the average strain
at the compression edge (strain between Points E and F).
The plastic hinge lengths of these two specimens, which will
be evaluated later as 2.5 times the wall thickness (300 mm
[11.81 in.]), are similar to the length between Points E and F
(400 mm [1.3 ft]). The strains at the ultimate drifts (the black
circles in the figures) were approximately 0.008, which
agrees with the ultimate strain of concrete eu computed in
the following considering the confinement effect.
FLEXURAL DRIFT CAPACITY
Simplification of plastic deformation
In this paper, flexural drift capacity is decomposed into
elastic and plastic components (Ru = Ry + Rp), as shown in
Fig. 16(a). The curvature at yielding fy is computed based
on the yield strain of longitudinal reinforcement (Fig. 16(b)).
fy =

ey
dc

(2)

where ey is yield strain of reinforcement; d is effective depth,


defined as the distance between the compression edge and
the center of the boundary column; and c is the neutral axis
depth computed from Eq. (3).
c=

Fig. 15Load and strain versus flexural drift of Specimen PM3.


tigate the effect of a perpendicular wall on flexural drift
capacity. The difference between these two specimens is
the existence of a perpendicular wall. The drift capacities of
NM3 (0.57%) and PM3 (0.61%) were similar. The compressive failure of the perpendicular wall of PM3 occurred before
the lateral strength degradation started. Because the perpendicular wall is provided with no confinement (Fig. 1), the
100

C
0.85b1 fct

(3)

where C is the compressive force of concrete computed


from Eq. (1); b1 is the reduction factor to determine the
neutral axis; fc is the compressive strength of concrete; and
t is wall thickness.
Theoretically, Eq. (3) is effective at the ultimate state
and ineffective at yielding; however, this difference may be
negligible because the amount of the wall reinforcements is
much smaller than that in the boundary column.
In this study, linear distribution was used for elastic curvature (Fig. 16(a)). Therefore, the elastic drift Ry is computed
from Eq. (4).
Ry =

D fy

h h2
= fy
h
2 6a

(4)

ACI Structural Journal/January-February 2013

Fig. 18Simplified model for plastic deformation.


Fig. 17Measured strain distribution of Specimen NM4.
(Note: NA is neutral axis.)
where Dfy is flexural displacement at yielding; a is shear span
length; and h is the specimens clear height.
The ultimate curvature fu is computed based on the ultimate strain of concrete (Fig. 16(c)), where eu is the ultimate
compressive strain of concrete defined in a later section
fu =

eu
c

(5)

The plastic drift is computed using plastic hinge length Lp.

R p = L p fu f y

(6)

Substituting Eq. (2) and (5) into Eq. (6) leads to the equation to compute the plastic drift.
Lp
c

Rp =
eu
ey
c
d c

c
ey
dc

ACI Structural Journal/January-February 2013

where the second term is the strain caused by the elastic


deformation. In this paper, ep is called the plastic component
of ultimate strain. Note that the rigid area in Fig. 18 rotates
Rp =

Lp
c

ep

(9)

(7)

Figure 17 shows the strain distribution measured using


LVDTs in Fig. 9 along the clear height of Specimen NM4
when the lateral force decreased to 80% of the maximum
strength. On the compressive side (the right edge), the
strain localized between Points E and F, whereas the strain
between Points F and G or G and H was quite small. On the
tensile side (the left edge), even the strain between Points C
and D exceeded the yield strain (0.0029 > ey = 0.002). It
is concluded that, for plastic deformation, the compressed
area is limited near the bottom of the wall, as indicated by
the shaded rectangle (compressed area) in Fig. 17 and 18,
whereas the area in tension is trapezoidal. The hatched area
in Fig. 18 is assumed as rigid. The strain at the compressed
edge in Fig. 18 is assumed to be uniformly ep, which equals
the term inside the parentheses of Eq. (7)
e p = eu

Fig. 19Assumed stress-strain model for concrete of


Specimen NS3.

(8)

around the neutral axis. Therefore, the height of the


compressed area in Fig. 18 can be regarded as the plastic
hinge length Lp. The deformation shown in Fig. 18 agrees
with the observed crack patterns and is similar to that
proposed by Hiraishi.14
There are two unknown parametersLp and epin
Eq. (9). In the following sections, these parameters are
examined using the tested specimens. Specimens tested by
Wallace and Orakcal2 and Tabata et al.5 are used because
flexural deformations of specimens are reported.
Plastic component of ultimate strain
The broken lines in Fig. 19 show the stress-strain relationships of confined and unconfined concrete in Specimen NS3
evaluated by the Saatcioglu and Razvi15 model, which is
applicable to rectangular sections. Figure 20(a) shows
the boundary element of NS3, where the shaded zone is
assumed to be confined. The confining pressure on the
shaded zone in the x-direction is computed assuming that
the horizontal bars are 100% effective. The confining
pressure in the y-direction is computed assuming that the
101

The ultimate concrete strain eu is defined as the strain


when the average stress of concrete decreases to 80% of the
maximum strength. The ultimate strain of Specimen NM2
without crossties is 0.0066, while the ultimate strains of the
other specimens are between 0.0078 and 0.0084. The ultimate strains of the specimens of Wallace and Orakcal2 are
approximately 0.008, except for Specimen TW2, which had
a strain of 0.0112. Figure 20(c) shows the boundary elements
of Specimens No. 2 and 3 of Tabata et al.5; although the
confinement ratio is higher than that of Fig. 20(a), its ultimate strain is 0.008 because its concrete strength was high
(70 MPa [10.15 ksi]). The strain at the compressive edge
when tensile reinforcement bars yield shown in Fig. 16(b)
(the second term in Eq. (8)) is approximately 0.001 in most
specimens. Therefore, ep in Eq. (8) is approximately 0.008
0.001 = 0.007 for all specimens except NM2 and TW2.

Fig. 20Assumed confined regions. (Note: 1 mm = 0.039 in.)

Fig. 21Relationship between Rp and lw/c.


crossties are 100% effective, while the cap bars at the wall
end with 90-degree hooks are 50% effective because the
observed strain of the cap bars was approximately one-half
of the yield strain. Figure 20(b) shows the boundary element
of Specimen NM2. Although only cap bars are provided in
NM2, the shaded area is again assumed to be confined. The
confining pressure in the y-direction is computed assuming
that the cap bars are 50% effective.
The solid line in Fig. 19 shows the average stress-strain
relationship of the boundary element calculated as the
weighted average of confined and unconfined concrete.
s = sc

tc
t tc
+ su
t
t

(10)

where sc is the stress of confined concrete; tc is the center-to


center distance between the horizontal wall bars (Fig. 20(a));
and su is the stress of unconfined concrete.
102

Plastic hinge length


As discussed in the Introduction, Wallace and Orakcal2 and
Tabata et al.5 assumed that Lp is equal to 0.5lw and 0.3lw,
respectively, where lw is wall length. If Lp is proportional
to lw, based on Eq. (9), Rp must be proportional to lw/c
because ep is similar for most specimens. The relationship
between Rp and lw/c is examined in Fig. 21. The variable
Rp is the plastic drift, which is the flexural drift minus the
elastic drift Ry computed by Eq. (4). The broken line in
Fig. 21 is the regression line, which is imposed to pass the
origin. The correlation coefficient is 0.70, where the results
of Specimens TW2 and NM2 are neglected because the
ep values of these specimens are very different from those
of the other specimens. The solid lines show the assumptions of Lp = 0.5lw and 0.3lw with ep = 0.007. They do not
agree with the regression line. The circles in Fig. 21 show
two of the data, which have different trends from the other
specimens. Specimen NS3, whose lw/t ratio is the smallest
(=8.5), exhibited a drift capacity twice that expected by the
regression line. On the other hand, Specimen NL2, whose
lw/t ratio is the largest (=20), exhibited a drift capacity 60%
of that expected by the regression line. Similarly, specimens
with small or large lw/t ratios are located above or below the
regression line, respectively. This tendency indicates that
hinge length is not simply proportional to wall length.
ACI 318-081 requires that the special boundary region
shall be longer than a/4, where a is shear span length. This
requirement implies that Lp in Fig. 18 equals a/4. To investigate whether the hinge length is related to shear span length
a, the relationship between Rp and the a/c ratio is shown in
Fig. 22. The solid line shows the assumption of Lp = a/4 with
ep = 0.007, which does not agree with the regression line
(the broken line). The correlation coefficient is 0.84 and
is better than that in Fig. 21. However, it is noted that the
results of Specimens No. 2 and 3 of Reference 5, whose a/t
ratios (=50) are much larger than those of the other specimens (=18 to 28), are located quite lower than the regression
line. This tendency again indicates that hinge length is not
simply proportional to shear span length.
Markeset and Hillerborg16 conducted uniaxial compression tests of plain concrete prisms with various lengths and
sectional dimensions. They observed that compressive failure
is quite limited within a certain length. They concluded that
the failure length was 2.5 times the shorter side length of
the compressed section (Fig. 23). In this study, wall thickness t is shorter than neutral axis depth c in all specimens.
Figure 24 shows the damage of Specimen NS3 at a 2% drift
ACI Structural Journal/January-February 2013

Fig. 24Observed failure of Specimen NS3 at 2% drift.


Fig. 22Relationship between Rp and a/c.

Fig. 23Compression localization.16

Fig. 25Relationship between Rp and t/c.

level (right after the strength degradation). The damage length


seems to be almost 2.5 times the wall thickness. Therefore,
2.5t is used for plastic hinge length Lp in this study. It should
also be noted that even in Specimen NM2 with the largest
thickness (t = 140 mm [5.5 in.]), the damage of concrete was
limited within the height of confinement (400 mm [1.3 ft]).
Therefore, it is concluded that the height of confinement
may be limited to 3t if the expected compressive strain is not
greater than 0.008.
Figure 25 shows the relation between Rp and t/c. The correlation coefficient is 0.94 and larger than the correlation coefficients in Fig. 21 and 22 (0.70 and 0.84). The broken line
in Fig. 25 shows the regression line. The solid line shows
Rp estimated from Lp = 2.5t and ep = 0.007. The solid line
reasonably agrees with the regression line, which implies
that the assumption of Lp = 2.5t is appropriate.
Figure 26 compares the observed and estimated flexural drift capacities (Ru = Ry + Rp). All test data except
Specimen TW2 are within 30% from the estimated drift
capacities. The observed drift of TW2 is much larger than
the estimated value. The observed compressive failure
zone of TW2 was also much wider than Lp = 2.5t. Similar

Fig. 26Comparison between estimated and observed flexural drift capacities.

ACI Structural Journal/January-February 2013

103

tendencies are observed for the test results of Paulay and


Priestley,8 which are not plotted in Fig. 26 because their
flexural drift capacities are not reported. These discrepancies
may be attributable to the difference of confinement. Recall
that the strain localization depicted in Fig. 23 is observed in
plain concrete. In the case of well-confined concrete columns
subjected to uniaxial compression, large plastic strain shall
be distributed uniformly over its entire length until strength
degradation starts (emax in Fig. 19). If the concrete column is
long enough, buckling occurs17 before the strain reaches emax.
The strain at buckling depends on the length of the column
and the tangential stiffness at the strain.17 The boundary
elements of Specimen TW2 and the specimens of Paulay and
Priestley8 almost satisfied the requirements of ACI 318-08.1 If
such a wall would be subjected to pure bending, uniform
curvature corresponding to emax or less would be observed
in its clear height when out-of-plane buckling would occur.
Therefore, for such a wall, eu in Eq. (7) should be replaced
with the strain at buckling and Lp should be a function of
shear span length, which would be much longer than 2.5t.
On the other hand, the cross-sectional areas of the horizontal bars and the crossties of the specimens tested by the
authors are 40% to 52% and 6% to 31% of those required by
ACI 318-08,1 respectively. The cross-sectional areas of the
confining bars of Specimens No. 2 and 3 tested by Tabata
et al.5 and Specimens RW1 and RW2 tested by Wallace and
Orakcal2 are 24 to 63% of those required by ACI 318-08.1 It
is concluded that Lp = 2.5t is valid if the confinement of the
boundary element is less than half of that required by the
seismic provisions of ACI 318-08.1 Otherwise, the equation
tends to underestimate the capacity.
CONCLUSIONS
The test results of 10 RC walls are described in this
study. Based on the experimental results and analytical
work presented in this paper, the following conclusions
are obtained:
1. All tested RC walls with limited confinement in the
boundary element failed in compression after flexural
yielding. The observed flexural drift capacity was between
0.4 and 1.2%.
2. Shear drift caused by lateral slip along the flexural crack
may be large if the ratio of the maximum shear force to the
compressive force of concrete is large (Fig. 13).
3. Plastic components of flexural drift can be modeled as
shown in Fig. 18, where the length of the compression zone
Lp is 2.5 times the wall thickness (Lp = 2.5t) if the depth
of the neutral axis is longer than the wall thickness and the
confinement of the boundary element is half of that required
by the seismic provisions of ACI 318-08.1
4. Ultimate flexural drift is defined as the drift when the
lateral force decreases to 80% of maximum lateral strength.
Ultimate flexural drift can be computed as the sum of
Eq. (4) and (7), where eu shall be determined as shown in
Fig. 19 considering the confinement effect.

104

5. The detailing of the boundary element shown in Fig. 2(c)


with the height of the confined area 3t is sufficient to obtain
an ultimate strain of eu = 0.008.
6. The effect of a perpendicular wall on ultimate drift is
negligible if the wall is not confined.
ACKNOWLEDGMENTS

This study was financially supported by the Ministry of Land and Transportation. The authors thank J. Wallace of the University of California,
Los Angeles (UCLA), who independently conceived the idea that hinge
length may be proportional to wall thickness, for valuable discussions. Data
provided by T. Tabata, H. Nishihara, and H. Suzuki of Ando Corporation
are greatly appreciated. The authors also thank H. Sezen of The Ohio State
University for critically reading the manuscript.

REFERENCES

1. ACI Committee 318, Building Code Requirements for Structural


Concrete (ACI 318-08) and Commentary, American Concrete Institute,
Farmington Hills, MI, 2008, 473 pp.
2. Wallace, J. W., and Orakcal, K., ACI 318-99 Provisions for Seismic
Design of Structural Walls, ACI Structural Journal, V. 99, No. 4, July-Aug.
2002, pp. 499-508.
3. Thomsen, J. H. IV, and Wallace, J. W., Displacement-Based Design of
Slender Reinforced Concrete Structural WallsExperimental Verification,
Journal of Structural Engineering, ASCE, V. 130, No. 4, 2004, pp. 618-630.
4. Design Guidelines for High-Rise RC Frame-Wall Structures,
National Institute for Land and Infrastructure Management, Kaibundo,
Tokyo, Japan, 2003, pp. 53-85.
5. Tabata, T.; Nishihara, H.; and Suzuki, H., Ductility of Reinforced
Concrete Shear Walls without Column Shape, Proceedings of the Japan
Concrete Institute, V. 25, No. 2, 2003, pp. 625-630. (in Japanese)
6. Kabeyasawa, T.; Shiohara, H.; Otani, S.; and Aoyama, H., Analysis of
the Full-Scale Seven-Story Reinforced Concrete Test Structure, Journal of
the Faculty of Engineering, V. 37, No. 2, 1983, pp. 431-478.
7. Orakcal, K., and Wallace, J. W., Flexural Modeling of Reinforced
Concrete WallsExperimental Verification, ACI Structural Journal,
V. 103, No. 2, Mar.-Apr. 2006, pp. 196-206.
8. Paulay, T., and Priestley, M. J. N., Stability of Ductile Structural
Walls, ACI Structural Journal, V. 90, No. 4, July-Aug. 1993, pp. 385-392.
9. Takahashi, S.; Yoshida, K.; Ichinose, T.; Sanada, Y.; Matsumoto, K.;
Fukuyama, H.; and Suwada, H., Flexural Deformation Capacity of RC
Shear Walls without Column on Compressive Side, Journal of Structural
and Construction Engineering: Transactions of AIJ, V. 76, No. 660, 2011,
pp. 371-377. (in Japanese)
10. Takahashi, S., Modeling for RC Shear Walls with or without
Boundary Column, PhD thesis, Department of Civil Engineering, Nagoya
Institute of Technology, Nagoya, Japan, 2011. (in Japanese)
11. Nagoya Institute of Technology, Flexural Deformation of RC Wall
with One Side Column, http://kitten.ace.nitech.ac.jp/ichilab/research/flexuralwall2010. (last accessed Nov. 5, 2012)
12. National Earthquake Hazards Reduction Program, Executive
Summary on Chile Earthquake Reconnaissance Meeting, NEHARP
Library, http://www.nehrp.gov/library/ChileMeeting.htm. (last accessed
Nov. 5, 2012)
13. Architectural Institute of Japan, AIJ Standard for Structural Calculation of Reinforced Concrete Structures, Maruzen, Tokyo, Japan, 2010,
pp. 484-491.
14. Hiraishi, H., Evaluation of Shear and Flexural Deformations of
Flexural Type Shear Walls, Bulletin of the New Zealand National Society
for Earthquake Engineering, V. 17, No. 2, 1984, pp. 135-144.
15. Saatcioglu, M., and Razvi, S. R., Strength and Ductility of Confined
Concrete, Journal of Structural Engineering, ASCE, V. 118, No. 6, 1992,
pp. 1590-1607.
16. Markeset, G., and Hillerborg, A., Softening of Concrete in Compression Localization and Size Effects, Cement and Concrete Research, V. 25,
No. 4, 1995, pp. 702-708.
17. Shanley, F. R., Inelastic Column Theory, Journal of the Aeronautical Sciences, V. 14, No. 5, 1947, pp. 261-268.

ACI Structural Journal/January-February 2013

Вам также может понравиться