You are on page 1of 11

Engineering Structures 56 (2013) 13351345

Contents lists available at ScienceDirect

Engineering Structures
journal homepage: www.elsevier.com/locate/engstruct

Fundamental principles of the reinforced concrete design code changes


in Chile following the Mw 8.8 earthquake in 2010
Leonardo M. Massone
Department of Civil Engineering, University of Chile, Blanco Encalada 2002, Santiago, Chile

a r t i c l e

i n f o

Article history:
Received 15 November 2012
Revised 2 July 2013
Accepted 8 July 2013

Keywords:
Shear wall
Concrete
Earthquake
Chile
Design
Connement
Buckling

a b s t r a c t
The Mw 8.8 earthquake in 2010 in Chile provided valuable information regarding the damage and potential design code changes for reinforced concrete (RC) structures. Many modern RC buildings suffered
severe damage, mainly in the form of concrete cover spalling, followed by longitudinal boundary bar
buckling and concrete crushing. The absence of wall boundary detailing explains such behavior, but older
structures and thousands of other buildings suffered minor or no damage at all. An analysis of some
potential factors that inuenced the damage caused is carried out and the fundamental principles for
the Chilean RC design code changes are exposed. Currently, a displacement-based approach is used for
detailing of the wall boundary elements as well as for establishing damage limitation. Compressive concrete strain cannot exceed 0.008, limiting indirectly the axial load, which is one of the potential precursors of the damage. Transversal reinforcement is also provided in zones of potential yielding of the
longitudinal reinforcement such that buckling due to compression preceded by tension is minimized; this
buckling is more likely in asymmetric cross-sections, such as T-shaped sections, where usually the web
goes into larger tensile and compressive strains given the variation of location of the neutral axis for the
web under tension or compression.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
The Mw 8.8 2010 earthquake in Chile impacted the country
over a coastal length of about 500 km, covering the most densely
populated regions. Within that region, the reinforced concrete
(RC) building stock increased in the last two decades, with more
than 1000 buildings with 1030 stories. During that period, the
construction of residential buildings greater than 15 stories became more common whereas earlier, construction was focused
on buildings with less than 15 stories. Modern buildings also presented a few characteristics that differ from older structures, such
as thin walls and discontinuities, among others [1]. The damage
was concentrated on newer structures, which also probably presents the largest building stock, specically taller buildings (usually over 15 stories). Most buildings presented no structural
damage or suffered very moderate damage, which required no
eviction. Some other buildings (40) presented severe damage that
was either repaired or, in rare cases, demolished. It is interesting to
note that intermediate situations were uncommon-that is, they
either presented serious damage or almost no damage at all, implying a brittle nature of the behavior. The severity of damage in modern buildings prompted the engineering community to study the
general nature of damage caused and propose an emergency code
Tel.: +56 229784984.
E-mail address: lmassone@ing.uchile.cl
0141-0296/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.engstruct.2013.07.013

to prevent current buildings from suffering similar damage in future earthquakes. Studies are currently being carried out to provide
additional information for future code considerations.
Modern RC building construction in Chile is based on shear
walls, which are designed according to ACI 318. The formal reference to ACI 318 was originally implemented though the Chilean
seismic code (NCh433.Of96 [2]), which requires the use of
ACI318-95 for RC building construction. However, based on the
good performance of RC shear wall buildings in Via del Mar in
the 1985 earthquake, the ACI 318-95 requirements for special
transverse reinforcement at wall boundaries to conne the concrete and restrain rebar bucking were eliminated. Seismic design
approaches for most RC buildings include modal response spectrum analysis, and the design spectra are based on historical
earthquakes.
As summarized by Massone et al. [1], typical buildings in Chile
have a large number of shear walls, with a total wall area over wall
plan area of roughly 3% in each principal direction of a building.
Although the wall area has been maintained over the years, newer
structures increased the number of stories, moving the average
building of 15 stories to about 25 oors. The increment of stories
has resulted in higher axial loads in newer buildings [3]. The design
code did not put a limit on the level of axial stress allowed for gravity load or combined gravity and lateral loads. The construction of
new buildings is also characterized by the use of a common layout:
longitudinal walls on each side of a central corridor that form a

1336

L.M. Massone / Engineering Structures 56 (2013) 13351345

spine and perpendicular (transverse) walls, or ribs, spaced at


roughly 5 m along the corridor. Thus, the walls in each principal
direction are often connected to form T-shaped or L-shaped
cross-sections.
A combination of large axial loads, relatively large lateral wall
displacement due to the earthquake, and poor wall boundary
detailing, besides wall discontinuities at the failure zone, resulted
in the damage of walls.
2. Research signicance
The Mw 8.8 earthquake in 2010 in Chile revealed that most
buildings in Chile were capable of sustaining a strong earthquake.
However, modern RC building construction was severely damaged
in some cases. This work presents an analysis of the damage
caused by actions of axial load combined with bending that characterized such buildings and exposes the fundamental principles
for the new RC design code, intended to provide safer structures
for future strong quakes. The main changes focus on special wall
boundary detailing, concrete damage limitation, and buckling
restraining (in cases not controlled by concrete compression
strain).
3. Damage in new buildings
Damage in buildings, located at the ground level or at the basement (Fig. 1), was usually observed in several rectangular or Tshaped shear walls and often over the entire wall length. In cases
where a long corridor was structured by T-shaped walls, seismic
movement in the short direction of the building caused signicant
damage, specically at the most compressed end of the web. In
general, spalling of concrete followed by concrete crushing and
the buckling of vertical reinforcement were observed. Typically,
the damage was concentrated over a short height, apparently because the buckling of vertical bars led to a concentration of the
damage [4], indistinctly in the lower or upper part of the story.
Commonly, the only weak constraints of longitudinal reinforcement against buckling were the large spacing of horizontal web
reinforcements (20 cm) terminated as 90 hooks with an L- or Ushape (Fig. 2). Fig. 2b shows, with double red lines, the L-shape
(left) and U-shape (right) termination of horizontal web reinforcement bars after being opened due to buckling of the longitudinal
reinforcement. Because the walls were thin, typically 1520 cm
thick, any spalling of cover concrete (about 2 cm on each side) resulted in an important reduction in the wall thickness, along with

Fig. 1. Typical shear wall damage.

the capability to sustain an important axial load, given the presence of basically unconned concrete. Upon spalling of the concrete cover, 90 hooks on the transverse reinforcement opened
due to buckling of the longitudinal reinforcement. Buckling of
the longitudinal reinforcement in walls was also combined, in
some cases, with the fracture of the reinforcement, likely due to
large cyclic strains.
Even though other failure types were observed, this work focuses on the code changes and their fundamental principles
regarding the failure caused by actions of axial load combined with
bending.

4. Construction characteristics and material response


The ineffective 90 end hooks of the horizontal web reinforcement imply that the concrete could be considered to be unconned
for practical purposes. Thus, for normal strength concrete, which is
the common case, peak compressive strength would be reached at
a strain of about 0.002. This also indicates that upon loss of concrete cover, rapid degradation is to be expected.
The unanchored horizontal reinforcement favors buckling of the
longitudinal reinforcement. In many cases, the damaged zone covers a height of about 400 mm or 500 mm (L), which is generally
about twice the horizontal reinforcement spacing, and also twice
the wall thickness. Considering that boundary longitudinal reinforcement is commonly 22 mm or 25 mm in diameter (d), the
buckled length to the bar diameter correlation is close to 20 (L/d)
[4]. If we only consider buckling degradation due to direct compression, the compressive strength is a function of L/d, as well as
the compressive strain attained, among other material parameters.
Fig. 3 shows the degradation of compressive capacity (strength ratio, fb/fs, at maximum compressive strain for symmetric cycles between the compressive capacity with and without buckling effect,
fb and fs, respectively) due to buckling for different L/d ratios and
different levels of maximum compressive strain (the total strain
would be twice that value). The curves were constructed based
on a model with 4 plastic hinges that has shown good correlations
with monotonic [5] and cyclic [6] response of reinforcements subjected to buckling. The model considers four lumped plastic hinges
located at the bar center (2) and ends (2). The resultant force and
moment are estimated by means of a sectional analysis that considers the nonlinear response of steel. The equilibrium is satised
in the deformed position, introducing geometric nonlinearity. For
the analysis, typical (not nominal) steel properties were considered
(yield stress-470 MPa, initiation of strain hardening-0.005, peak
stress of 707 MPa at 0.12 strain). In Fig. 2b, the ratio L/d is close
to 20, and considering that the transversal displacement is close
to 3d or 4d, then, based on shortening only due to the rotation of
the plastic hinges leading to transversal displacement, the total
strain from the straight position to the shown buckled position
would be close to 7% (this could be q
monotonic
compression
or after

2
2
an initial tension from a length of 2 10d 4d 21:5d, which
yields a total deformation of (21.5d  20d)/21.5d = 7%). This indicates that the buckled reinforcement would have almost no residual capacity (Fig. 3).
Fig. 2b shows a series of buckled bars on the right side of the
wall section, whereas on the left side last three bars also present
fracture at mid-height of the buckle length. Fracture in this case,
as can be seen in the gure, shows no indication of a reduced
cross-section or necking of the bar due to stress concentration at
the fracture zone, as commonly seen in direct tension tests. Failure
is rather associated to fracture due to fatigue, which presents an almost clean horizontal cut of the reinforcement. This type of failure
under low strain or stress is common in mechanical equipment
requiring millions or thousands of cycles. In our case, where few

L.M. Massone / Engineering Structures 56 (2013) 13351345

1337

Fig. 2. Boundary wall reinforcement (a) general view, and (b) fracture and buckling.

large wall top displacements (1.5%), for a stable response (before


degradation), where lw is the wall length. Thus, for a typical wall
length of 5 m, the expected plastic hinge length is about 1.7
2.5 m. That is about 35 times what was observed in wall boundaries without detailing. If we consider that most of the strains in
the longitudinal reinforcement in walls in Chile are inelastic (conservative estimate, since a large percentage distributes over the entire length of the wall) and concentrate within the plastic hinge
length, then after restraining buckling, the plastic hinge length
could increase in about 35 times, reducing the strain in about
the same magnitude. This would rapidly increase the bar life under
fatigue.

fb/fs (degradation)

L/d=6
L/d=10

0.8

L/d=14

0.6

0.4

0.2
0.01

0.02

0.03

0.04

Strain (max.)
Fig. 3. Compressive degradation due to buckling for different L/d and strain values.

cycles with large wall lateral displacement demands were observed from the earthquake associated with a large strain magnitude, fracture can occur in less than 100 cycles. This is
commonly known as low cycle fatigue. Tests by previous researchers have provided an expression that allows for an estimation of
half cycles to failure (2Nf) for specic total strain amplitude (ea-half
of the range amplitude). Brown and Kunnath [7] who studied cyclic
tests under symmetric tensioncompression determined that for
25 mm-diameter bar, which is a typical bar diameter for boundary
bars in Chile, the bar life fatigue is given by ea = 0,08(2Nf)0.36, from
where, given a 3.5% amplitude strain (assuming a total strain range
of 7%, as mentioned previously), the number of half cycles that results is 10 (i.e., 5 cycles), which is consistent with the number of
cycles of large magnitude that can be observed in the earthquake
records. That expression also indicates that in order to avoid low
cycle fatigue failure for less than 50 cycles, the total strain range
should not go over 3%.
The question that remains then is-how could we observe less
strain demands in longitudinal reinforcement in order to avoid rebar fracture?, or why was not observed fracture in all cases? One of
the important aspects that caused damage in walls is the impossibility of the nonlinear or plastic deformation to distribute over a
height larger than 400500 mm. That damaged zone is consistent
with the buckled length of the reinforcement. Distribution of damage (or plastic deformation) over a larger zone would reduce the
curvature demands, and therefore the compressive strain demands
for a prescribed wall top displacement. Test results from the literature (e.g., Thomsen and Wallace [8]) have shown that the plastic
hinge could easily be in the range of 0.33lw and 0.5lw for relatively

5. Axial load and bending in slender wall tests


The following section describes ndings associated to slender
wall tests and analysis from the literature, which particularly deals
with connement quantity, reinforcement buckling restraint, impact of wall section asymmetry in strain history, among others.
These aspects are relevant in the response of slender walls, which
also explain the damage observed in walls in Chile earthquake, and
therefore should be incorporated in their design.
Tests on slender walls have commonly been carried out providing enough shear strength in order to promote a failure that combines bending and axial actions named as exo-compression.
Observed damage in walls from test programs show similar characteristics as the walls damaged during the 2010 earthquake, such
as spalling of concrete in the most compressed area, followed by
buckling of the longitudinal reinforcement and fracture after a
few cycles [8,9]. The tests by Thomsen and Wallace [8] were carried out in cantilever with a lateral point load at the wall top,
resulting in a shear-span to depth ratio of 3.1, where the response
was controlled by exure. Shear lateral deformations at the top
were estimated to be about 10% of the total displacement [10].
An axial load of approximately 0.1fc0 Ag , where fc0 corresponds to
the compressive strength of concrete and Ag the wall cross-section
area, was also applied to the specimens. Four specimens were
tested, two rectangular walls (RW1, RW2) and two T-shaped walls
(TW1, TW2). Rectangular walls were 1.22 m long and 0.1 m wide.
T-shaped walls also had a 1.22 m long and 0.1 m wide ange, as
well. The main difference between TW1 and TW2 was the amount
of connement, tighter and more prolonged over the web boundary in TW2 [8]. In the tests, although the failure was similar in
all cases, in the specimens with lower amount of connement,
more specically, those with larger transverse reinforcement spacing, the buckling of longitudinal reinforcement led to damage
localized in a narrower section, which covers a section larger than

1338

L.M. Massone / Engineering Structures 56 (2013) 13351345

the horizontal web reinforcement spacing (Fig. 4a, TW1). On the


other hand, in specimens with a tighter spacing of transversal reinforcement at wall boundaries, damage spreads over a higher section (Fig. 4b, TW2).
The connement effect was also studied in T-shaped (TW) and
rectangular (RW) wall cross-sections. In both cases, a larger
amount of connement resulted in larger displacement capacity.
It is important to mention that in this test program, one of the
parameters that were changed to modify the connement quantity
was the spacing of the transversal reinforcement. Therefore, the
larger displacement capacity is not only affected by the connement level, but also by the tendency of buckling of the longitudinal
reinforcement.
The cyclic lateral top displacement of the walls also inuences
the wall behavior. The asymmetry in the wall cross-section or reinforcement should be taken in consideration when analyzing the
cyclic response of walls. Fig. 5 shows the strain history for the extreme bers of a ber model with material nonlinearity that characterizes the response of TW2 [8], which have been validated
against the experimental data [11]. The response of an extreme ber for RW2 is also included for selected drift levels. Cycles with
similar ber strain represent overall cycles with identical wall
top displacement (commonly 2 full cycles). Upon reversal, the
top displacement was set to the same value.

The gure presents important characteristics regarding design


and detailing of walls that need to be considered or at least understood, especially when they are non-symmetric. For identical wall
top displacements for 2 different loading directions (e.g., phase I
and II in Fig. 5), the maximum compressive and tensile strain values at both wall ends are vastly different. When compressing the
ange (phase I), the longitudinal reinforcement at the web goes
in tension with a neutral axis depth usually within the ange zone.
In order to reach relatively large top displacements, and therefore
large curvatures, large tensile strains are necessary in the web end
(A), and slight compressive strains in the ange end (B). Upon
reversal (phase II), the narrower section of the web and the relatively large amount of reinforcement in the ange area force the
neutral axis depth to move away from the web end resulting in
more damage in concrete under compression (crushing) at the
web end (A) compared to the ange end (B). This is consistent with
the experimental evidence in test programs with T-shaped [8] and
L-shaped walls [9], as well as ndings after the 2010 earthquake
[4]. The response of RW2, which is a symmetric rectangular section, shows smaller tensile and compressive strain values than
the web end (A) of TW2, indicating that under similar connement
and bar buckling constrain is more likely to degrade the response
of the T-shaped section for similar drift levels. Tensile strains in

Fig. 4. Damage at 2% drift (nominal) (a) TW1 and (b) TW2.

Strain (extreme ends)

0.04

A - TW2
B - TW2
RW2

0.02

A
0

-0.02

+ tension
- compression

II

1%

Phase I
1.5%

Drift Level
Fig. 5. Strain history for extreme bers (ber model)-TW2 and RW2.

Phase II

1339

L.M. Massone / Engineering Structures 56 (2013) 13351345

RW2 are larger than tensile strains in the ange (B), but with similar compressive strains.
Concrete crushing might not be the only concern about the
behavior of non-symmetric sections. Large tensile strains in the
web end (A) followed (upon reversal) by large compressive strains
in the same point can lead to buckling of the reinforcement, making T-shaped walls (or in general non-symmetric) more prone to
this type of damage. Moreover, even large tensile strains followed
by little or no compressive strains can lead to buckling [1214], as
shown in Fig. 6. Fig. 6 shows the analytical response (plastic hinge
model [5]-as previously described for L/d = 14) of a bar subjected to
buckling for two different loading histories, but with the same total
strain variation (6%, Fig. 6a ranging from 3% to 3%, and Fig. 6b
ranging from 2% to 4%). In both cases, onset of buckling (marked
in the gure) is observed almost immediately upon reversal from
tension.

an inelastic component that accounts for the additional curvature


once the ultimate curvature has been reached (/u). The inelastic
curvature usually increases linearly within the lower portion of
the wall, starting from the point where yielding is observed in
the boundary longitudinal reinforcement [17,18]. Although the
curvature tends to increase linearly, simplied models [19] have
characterized the inelastic component as an equivalent rectangular
distribution of curvature of length lp, which represents the plastic
hinge length. Thus, the wall top lateral displacement (du) can be
determined using Eq. (1), if a triangular distribution of the forces
is applied along the wall height (hw). For simplicity, the ACI 318
code adopted a modied version of Eq. (1). Instead of considering
the elastic and inelastic components, only one component is considered, which accounts for the entire ultimate curvature at the
wall bottom, locating the plastic hinge length at the wall base
and neglecting the elastic component in the upper section (Eq.
(2), Fig. 7b).

6. Displacement model for detailing

du

The current design procedure for wall boundary detailing in ACI


318-08 [15] is displacement-based [16]. ACI 318-08 uses a simplied compatibility model to determine the maximum compressive
strains in concrete due to wall top displacement, requiring special
boundary connement when the compressive strains are larger
than 0.003. The procedure assumes a cantilever wall with lateral
loads either concentrated or distributed over the height of the wall.
Fig. 7 shows a representative wall with a triangular distribution of
lateral loads. The traditional curvature distribution model assumes
an elastic component until the yield curvature is reached (/y) and

du u lp hw

-0.02 to 0.04

(b)

(a)

0
-400

Buckling
No Buckling

Buckling
No Buckling

-800
-0.04

-0.02

0.02

0.04

-0.02

0.02

Strain

Strain

Fig. 6. Initiation of buckling (a) 3% to 3% strain range, and (b) 2% to 4% strain range.

The simplied formulation assumes that the elastic component of


the deformation can be indirectly included in the inelastic portion
such that all deformations are concentrated in only one sectionthe plastic hinge. This simplication implies that, given a wall top
displacement value, the required curvature depends on the top displacement, wall height, and plastic hinge length (Eq. (2)). That is, in
Fig. 5, the maximum curvature magnitude should be equal, for

-0.03 to 0.03

800
400



11
lp
2
y hw u  y lp hw 
40
2

Fig. 7. Cantilever wall model for boundary detailing (after [16]).

0.04

1340

L.M. Massone / Engineering Structures 56 (2013) 13351345

example, for loading phases I and II, considering that top displacements are identical. Curvature can be estimated from the gure as
the difference between the strain values at extreme ends (distance
between peaks values of A and B-shown in Fig. 5 with vertical arrows) over the length of the wall. As can be seen, in phase I, the curvature is about 25% larger than in phase II. That difference is larger
with a smaller top displacement and smaller with a larger top displacement. This comes from the fact that the yield curvature, and
therefore the yield top displacement for T-shaped walls are different in each direction. The yield curvature tends to be smaller in
T-shaped walls with anges in compression [20], resulting in larger
curvature demands in the nonlinear range when compared to the
same top displacement for the wall with the ange in tension,
where the yield curvature is larger.
7. Axial load
Damage during the 2010 earthquake in Chile was more common in modern RC wall buildings, rather than old RC wall structures. Older RC buildings were constructed usually past the
1950s. Newer structures were usually constructed around year
2000, and featured narrow wall thickness (for taller buildings)
and the presence of discontinuities, which were less common in
previously built structures. Considering that almost no building
had connement, two main, different aspects arise: (1) axial load
level, and (2) discontinuities. Both can affect the strain distribution
(e.g., Eqs. (1) and (2)). The treatment of discontinuities is a specic
and difcult task and most design tools are strength-oriented, such
as strut-and-tie models [23]. For the purpose of this work all analysis of strain distribution are related to axial load.
Few research programs have focused their attention to the axial
load effect in slender walls, especially if we consider loads larger
than 0.2fc0 Ag [21,22]. The work by Zhang and Wang [21] included
a test program that compared mid-height (effective height = 1.5 m,
length = 0.7 m) RC and steel reinforced concrete (SRC, partial
boundary longitudinal reinforcement replaced by steel sections)
walls under cyclic lateral loads. The specimens were tested in cantilever with a point top lateral load and a constant axial load applied at the wall top. The RC walls (SW7 and SW8) had similar
geometry, reinforcement conguration, and material properties,
except for the boundary reinforcement ratio (SW7: qb = 0.062,
SW8: qb = 0.046) and axial load (SW7: 0.24fc0 Ag , SW8: 0.35fc0 Ag ).
Test results for the overall load (backbone curve) vs. displacement
(at 1.5 m from base) are shown in Fig. 8. The gure shows a larger
displacement capacity for the specimen with a lower axial load,
although it had a larger amount of longitudinal boundary reinforcement. It is also important to point out that with the prescribed

connement (Ash/sbc = 0.011, where Ash is the transversal stirrup


area in the direction of analysis, s is the stirrup spacing and bc
the distance between outer stirrup legs), the specimen with an axial load of 0.35fc0 Ag reached a drift larger than 1.3% (200 mm) before the initiation of signicant degradation. The length of the
conned zone (100 mm) is consistent with a design displacement
of 0.42% drift, according to Eq. (2) du =hw 0:003
: 700
). This assumes
250
2
that the connement ends at the location with a strain of 0.003
(100 mm from wall edge), which results in maximum compressive
strain of 0.0042 (0:003
 350) for a neutral axis depth of about lw/2
250
and lp = lw/2. This indicates that such expression and the connement supplied is conservative for small displacement demands. If
we consider the initiation of important degradation (200 mm)
and assume the same neutral axis depth (reasonable for stable
boundary element) the maximum compressive strain increases to
1.3% (1:3%  350= 700
), which is already in the degrading branch of
2
the compressive stress vs. strain response of conned concrete.
Failure was, in this case, promoted by out-of-plane buckling.
The axial load effect can be studied through a simple model, for
example, the one presented by Eq. (1). Considering that the yield
curvature can be estimated as 1.5ey/lw (ey = 0.0021 yield strain,
associated to a nominal yield stress of fy = 420 MPa), and the ulti, the lateral
mate curvature for unconned concrete as u 0:003
c
top drift capacity from Eq. (1) becomes, for lp = lw/2,





1:5ey
du
hw 12 0:003 lcw  1:5ey 1  4hlww , which is dependent
11
40
hw
lw
on c/lw and hw/lw, where c represents the neutral axis depth and
hw/lw the wall aspect ratio. Considering that c is determined by
equilibrium at the sectional level, the neutral axis depth depends
on the level of axial load, as well as of other factors such as geometry, steel quantity and material properties. For a common wall
length of 5 m and a story height of 2.5 m, the drift capacity ratio
is related to the number of stories (N = hw[m]/2.5) and the c/lw ratio. The drift capacity vs. the number of stories is shown in Fig. 9 for
c/lw values of 0.1, 0.2, 0.4 and also for yielding (balance condition)
of the boundary reinforcement with solid lines. Additionally, for
comparison purposes, a dashed line that allows compressive
strains reaching a 0.008 value is considered for c/lw = 0.4.This is
consistent, as explained later, with the new Chilean wall design
limitation to compressive strain. In this case, it is assumed that
the concrete is conned in order to reach such values without signicant degradation. Other values of c/lw are not shown since they
reach drift values larger than 2%.
The spectral drift demand can be estimated using the elastic
spectral displacements for any ground record. Considering that
the interest is focused in relatively tall buildings, the equal displacement rule is considered to estimate the inelastic displacement, and an amplication factor of 1.3 is used to estimate the

250

0.02

0.015
150

u/hw

Load (kN)

200

100
SW7
50

0.01

0.005

SW8

Damaged
buildings
San Pedro record
EW dir., 2% damp.

0
0

10

20

30

Top (@1.5m) displacement (mm)


Fig. 8. Axial load effect [21].

40

0
5

10

15

20

N
Fig. 9. Axial load effect in drift capacity.

25

30

1341

L.M. Massone / Engineering Structures 56 (2013) 13351345

displacement demand in a multistory building. Due to the large


stiffness of Chilean buildings, the fundamental period based on uncracked stiffness (element gross area) is estimated as T = hw/70,
with hw the height of the building from the ground level in (m).
Most buildings have fundamental periods between hw/40 and hw/
140, according to a collection of analytical models by Caldern [3].
The fundamental period based on cracked stiffness is estimated
as Tcr = 1.5 T. Thus, the displacement demand that depends of the
period, and therefore of the height, is plot in Fig. 9 as drift demand
vs. number of stories. The San Pedro record [24] in the strongest
horizontal direction (EW), measured at about 5 km south-west of
downtown Concepcin, is shown for the analysis for a 2% damping.
As seen in Fig. 9, the larger the axial load, that is larger c/lw, the
lower the drift capacity. For c/lw = 0.2, the drift capacity is already
overcome by the drift demand, which occurs for buildings with a
number of stories around 15. This analysis is consistent with the
observations since most buildings that presented damage were in
the range of 1520 stories. Besides, that could explain why old
building with probably lower quality construction materials presented good responses. They commonly exhibit a lower axial load,
associated to low-rise buildings, with the range being usually under 15 stories. If we also consider that modern buildings presented
discontinuities, that would result in strain concentration amplifying its magnitude, and therefore, reducing the drift capacities.
The previous analysis assumes that the plastic hinge length is
lp = lw/2. The evidence from the past earthquake indicates that
the localization of damage as seen in Fig. 1 does not spread as suggested by the common expression due to concentration of damage
within a small section that usually exists between 1 or 2 lines of
the transversal web reinforcement, that is, around 400 mm. The
severity of the damage indicates that at such a stage, degradation
of the response is expected, and therefore, that should not be expected in walls before the degradation has set in. Experimental evidence indicates that before degradation, the values for plastic
hinge length are close to 0.5lw [16]. However, those walls had an
axial load of about 0.1fc0 Ag , which might not represent what was
observed in Chile, where the axial loads have been estimated to
be much higher than 0.1fc0 Ag for tall buildings [1].
Based on beam tests, it has become common practice to assume
that the plastic hinge length lp of a concrete wall varies from 0.5 to
1.0 times the larger horizontal dimension of the wall (lw). Thus,
0.5lw is used as a lower-bound estimate of lp to make a safe estimate of the displacement capacity from the curvature capacity.
Sawyer [25] developed an equation for plastic hinge length in
beams. In this work, it was assumed that the ratio of the yield moment to the maximum moment is 0.85 and it was determined that
the location of the initiation of longitudinal reinforcement yields at
a distance 0.15z from the maximum moment point, where z is the
distance from zero moment to the maximum moment. After that,
several expressions have incorporated both parameters: (1) one
associated to the length of the walls and (2) another associated

The observed damage from the earthquake as well as information gathered from acceleration records led the Chilean civil engineering community to demand changes to existing standards-the
design of reinforced concrete (NCh430.Of2008, [26]) and the seismic design of buildings (NCh433.Of96, [2]). The changes in this
seismic design code mainly focused on developing both new design displacement spectra and a modied soil classication, which
were reected in its replacement, the DS No. 61 [27], which had a
previous version earlier in 2011.
Regarding the modications of the design standard for reinforced concrete, it focused on requirements for an improved
behavior of reinforced concrete walls, whose replacement was
the DS No. 60 [28] (also a modication of an earlier version). The
code uses the ACI 318-08 as the base code and includes modications that focused on special wall design.

1.5

SoilD
SoilC
Soil B
Soil A

0.008

8. New reinforced concrete requirements

(a)

(b)
1

c/lw

1.3Sd (Tcr)/hw

0.012

to the height of the walls. These expressions recover the effect of


location of the yielding moment, extending their use to columns
and walls. Other parametric studies have been carried out that
incorporate other variables, such as the axial load [17,18]. The axial
load, as mentioned before, changes not only the overall response,
but at the sectional level, it also changes the post-yield stiffness
of the momentcurvature diagram. The smaller the axial load,
the larger the post-yield stiffness, given that damage and therefore
degradation, is postponed for larger curvature values. A larger positive post-yield stiffness, which results in a larger maximum moment, allows for a larger plastic hinge length at the maximum
moment since the yield moment would move upward in a cantilever wall. Bohl and Adebar [17] suggest a variation of the plastic
hinge length that changes with the axial load (P), resulting in
lp 0:2lw 0:05z1  1:5P=fc0 Ag . If we only consider the additional impact of the axial load, the larger the c/lw value, the smaller
the lp value observed, which results in a smaller drift capacity,
accentuating the problem with the axial load. This effect is not included since the c/lw value would depend on the axial load, as well
as steel quantity, wall geometry, and material properties, among
others. If we consider a rectangular symmetric cross-section without web reinforcement with normal strength concrete, where all
boundary reinforcement is yielding, the relationship between the
axial load and the neutral axis depth, by force equilibrium, reduces
to f 0PAg 0:72 lcw . In this case, a c/lw = 0.4 would reduce the plastic
c
hinge length by about half. If we consider a more common section
conguration, such as T-shaped wall (ange length similar to web
length) with distributed web reinforcement (qw = 0.0025) with the
ange under tension, the axial load normalized over the wall web
area is reduced only by about 20%. All this indicates that the axial
load not only reduces the drift capacity by increasing the neutral
axis depth, but also reduces the ability of the wall to distribute
plasticity over its height.

0.5

0.004

SoilD
SoilC
Soil B
Soil A

0
0

Tcr (s)

Tcr (s)

Fig. 10. (a) Drift demands, and (b) neutral axis depth limitation for zone 3.

1342

L.M. Massone / Engineering Structures 56 (2013) 13351345

40

Stress (MPa)

30

0.008
20

0.003

Confined
Unconfined

10

0
0

0.005

0.01

0.015

0.02

Strain
Fig. 11. Conned and unconned concrete.

One of the rst modications involves the axial load, limiting


the maximum axial load to 0.35fc0 Ag for every wall element at
any location. Considering that this limitation may not be suitable
for asymmetric sections, which may have premature compression
failure, a damage limitation requirement against brittle compressive failure was adopted. Based on the simple plastic hinge models,
it requires that when slender walls (hw/lw P 3) reach the building
design displacement (du), a compression strain in concrete of
0.008 in the extreme most compressed ber at the critical wall section is not exceeded. This strain (ec) can be estimated, derived from
Eq. (2), as ec u c h2dwulwc , after assuming that lp = lw/2. Alternatively, the elastic deformation capacity of the wall can be incorporated to reduce the requirement of inelastic displacement (Eq. (1)).
The explicit consideration for the plastic hinge length (lp = lw/2)
was only imposed for Eq. (2), since the strain estimates tend to
be conservative (elastic component neglected). As it was exposed

(a)

before, for larger axial loads, the plastic hinge reduces which could
be considered in Eq. (1). In both cases, it is assumed that there is a
potential plastic hinge where the inelastic deformations are concentrated, and therefore, a special detailing requirement for the
edge of the wall is provided. Considering that by exceeding a compressive strain of 0.003 in the most compressed ber in the wall
(ACI 318-08) special detailing (transversal reinforcement) needs
to be provided at wall edges, the consideration of compressive
strain limitation of 0.008 seeks to limit the potential damage that
may occur. This limitation indirectly restricts the axial load level
(through the neutral axis depth) in the walls, which is dependent
on the expected demand for the wall displacement. Thus, for buildings in seismic zones closer to the hypocenter and in soft soils, the
expected wall top displacement demand will be higher, resulting
in larger compressive strains and, thus, it will be more restrictive
with the axial load level. Since the limitation is not strictly on
the axial load, other wall characteristics can be modied (e.g.,
amount of longitudinal reinforcement, yielding of the reinforcement, and geometry).
In order to study the impact of the damage limitation requirement, the neutral axis depth limitation is estimated on the basis
of the expected drift demands provided by the DS No. 61 [27].
The drift demands are estimated for zone 3 (coastal-highest seismicity) and for the four soil types that provide spectral displacement (AD). A fundamental period for an uncracked lower bound
stiffness of T(s) = hw(m)/50 is selected. The cracked stiffness is estimated as Tcr = 1.5 T [27]. Drift demands are plotted in Fig. 10a,
which indicates that the maximum drift level yields a value of
about 1.1%. The low drift levels are promoted by the large building
stiffness. Knowledge of the drift demand and the use of Eq. (2) provide the limitation to the neutral axis depth for a compressive
strain of 0.008, as shown in Fig. 10b. Zone 3 soil D yields the smallest value (c/lw = 0.37), which is consistent with a low axial load. For
a symmetric rectangular section with yielding boundary reinforcement and no web reinforcement, it yields an axial load of P/
fc0 Ag = 0.27. The axial load reduces considerable in non-symmetric
sections such as T-shaped walls, yielding values that might be half

(b)

Fig. 12. Onset of buckling (a) rectangular, and (b) T-shaped cross-section.

L.M. Massone / Engineering Structures 56 (2013) 13351345

of the symmetric rectangular wall case, when normalized over the


entire wall cross-section area, since the term Ag would represent
the web area in that case. Thus, the code captures the less ductility
observed in non-symmetric sections by improving their displacement capacity with, for example, lower axial loads.
Given a prescribed wall top displacement or curvature, a compressive strain limitation forces larger tensile (commonly) strains
in the opposite side of the wall, which might induce tensile fracture of the longitudinal boundary reinforcement. Low amounts of
boundary longitudinal reinforcement and axial load reduces the
neutral axis depth favoring the same situation. In Chile, it is common to provide a steel quantity of 5% of the boundary area (0.5%
over wall area) at each boundary and no less than 3% in most cases
for modern buildings over 15 stories [29]. If we consider the extreme case of no axial load, for a rectangular section with symmetric boundary reinforcement (qb = 3%) and no web reinforcement,
where the compressive reinforcement might not be yielding, and
for simplicity under zero strain, due to the low neutral axis depth,
the neutral axis depth reduces, by equilibrium, to lcw
0:030:1t w fy 420 MPa
0:85b1 0:85fc0 25 MPat w

0:07, for a boundary length of 0.1lw, wall

thickness tw, and a Whitney stress block coefcient b1 = 0.85. From


Fig. 10a, for a maximum drift of 1.1%, Eq. (2) reduces to hdwu
0:011 u lp  eclwes

lw
,
2

which together with

c
lw

ecec es 0:07, yields

to es = 0.02, where es and ec are the extreme tensile and compressive strains in the wall. A 2% tensile strain is far from causing tensile fracture of the boundary reinforcement. A stable compressive
boundary element (connement and buckling restriction might
be required) capable of equilibrating the force from the tensile longitudinal reinforcement (and axial load) that guarantees a nearly
constant plastic hinge length (here assumed as lw/2) and low drift
demands (here assumed no more than 1.1%) will be capable of preventing tensile fracture of the reinforcement. Smaller plastic hinge
length (or a hinge that reduces its length with drift) or larger drift
demands would result in larger tensile strain values in the longitudinal reinforcement. Non-symmetric sections would probably
present larger neutral axis depth, reducing even further the tensile
strain.
Another important aspect related to the limitation of damage is
the fact that in order to have the distribution of plasticity along the
height of the wall, a positive post-yield stiffness is required. Considering that concrete in compression should be conned if the
compressive strain is larger than 0.003, low strain values in unconned concrete would present just slight degradation of capacity
associated with concrete in compression, but exhibit an overall stable response (e.g., momentcurvature). If we consider the connement requirements for wall special boundary elements by ACI 318f0

08, that is, Ash  0:09bc s fc0 , for normal material properties
y

(fc0 = 25 MPa and fy = 420 MPa), the connement pressure reduces


to f1

Ash fy
bc s

 2:25 MPa. Considering the connement concrete

model by Saatioglu and Razvi [30], and a wall length of 5 m under


an axial load consistent with c/lw = 0.2 (i.e., P/fc0 Ag = 0.14 for a symmetric rectangular section with yielding boundary reinforcement)
results in a conned length of 0.63 m (=0.250.005/0.008) for a
maximum compressive strain of 0.008 and connement over compressive strain values of 0.003. The connement effectiveness is
determined for connement reinforcement separated transversally
by 200 mm and longitudinally by 150 mm. The unconned concrete would present a peak stress at a strain value of about 0.002
in compression and 15% strength degradation at 0.0033. Thus,
the conned concrete would present a peak stress of 33 MPa at a
strain of 0.0053 with 15% strength degradation at a strain of
0.011. This indicates that the limitation of compressive strain of
0.008 corresponds to just the initiation of degradation of conned

1343

concrete, and therefore positive post-yield is expected at the lateral top drift demand, similarly to the case of unconned concrete
with maximum compressive strain below 0.003. The stress vs.
strain response in compression is shown in Fig. 11 for conned
and unconned concrete, where both the compressive strain limits
are shown (0.003 for unconned concrete and 0.008 for conned
concrete).
In cases requiring connement, the wall thickness must be at
least 300 mm, and the conned length may not be less than the
width of the wall in the conned zone to ensure proper containment of the transverse reinforcement. In this sense, and also to ensure good concrete placement and behavior, bar diameter size
limitations are imposed on both the longitudinal reinforcement
(db_long 6 1/9hSBE, with hSBE thickness of the wall at the boundary
element) and the transversal reinforcement (db_transv P 1/3db_long)
at wall boundary elements, as well the vertical (s 6 1/2hSBE) and
horizontal (hx 6 1/2hSBE, 200 mm) separation of supported vertical
bars by transversal reinforcement. Additionally, both the stirrups
as cross-ties must have their ends bent at an angle greater than
or equal to 135.
Where connement is not required, but there is a longitudinal
reinforcement ratio greater than 2.8/fy, and it is expected that this
reinforcement might yield, the spacing of the transverse reinforcement is limited to 6db and 200 mm, with db the diameter of the longitudinal reinforcement. This limitation is consistent to the one
provided in ACI 318-08; however, here the spacing limitation to
6db prevents or delays reinforcement buckling to wall drift levels
that are expected in stiff shear wall structures in Chile. Thus, for
example, in asymmetrical walls, such as T-shaped wall sections,
where the compression strain do not exceed 0.003, they are required to have buckling restriction as these bars do not reach signicant compression deformation, but they might buckle if they
have previously endured large tensile strains. They could even
buckle while still elongated, that is, under tensile strain, but subject to compression stress. In order to quantify the impact of such
limitation, the drift capacities should be estimated based on the
initiation of bar buckling, providing a limit state for the buckling
condition.
Collection of experimental data by Rodriguez et al. [12] for steel
coupons loaded under cyclic strain history, recalling bars in columns and beams under seismic actions, show how the onset of
buckling is dependent on stirrups spacing (xed-end condition).
The onset of buckling is dened experimentally, by Rodriguez
et al., as the point where the instability of the bar was captured
with extensometers place on two sides of the bar, where one of
the extensometers presented increase of compression (concave
side of the buckle bar) and the other decrease of compression (convex side of the buckle bar). Initiation of buckling was dened at the
point where the difference between both sensors was 20% of the
compression observed in the concave side. The strain at buckling
was studied by means of a strain ep* that corresponds to the total
strain from the previous zero stress point until buckling was determined. This, for simplicity, can be approximated to be the entire
strain between the maximum tensile strain and the subsequent
compressive (or tensile) strain at the onset of buckling. If we consider a rectangular wall cross-section with symmetric reinforcement under a seismic action from left to right that results in a
strain prole described in Fig. 12a (direction 1), the extreme bers
would reach strains et and ec at both ends. Upon reversal (direction
2) and considering that the same top displacement is applied, the
maximum strains would be the same, but in reverse order, that
is, ec and et. Thus, each extreme ber would have reached a maximum strain of et + ec, which is approximately ep* once buckling begins. According to Rodriguez et al., for stirrups spacing of s = 6db,
the onset of buckling occurs, in average, at about ep* = 0.04. If we

1344

L.M. Massone / Engineering Structures 56 (2013) 13351345

consider Eq. (2) (and lp = lw/2), then hdwu ecc

lw
2

eclwet

lw
2

 0:04
0:02,
2

which indicates that the reinforcement under such transversal


constraint results in drift capacity of 2%, which is larger than expected in RC shear wall buildings in Chile, according to Fig. 10a.
In an extreme case for nonsymmetrical cross-section (e.g., T wallFig. 12b), where the neutral axis depth for the web in tension falls
inside the ange, c  lw, if we consider the same situation (conservative) when the web is in compression, then hdwu ecc l2w
ec et lw
2lw
2

 0:04
0:01. The drift capacity reduces to 0.01, which is
4

still a considerable drift capacity, and similar to the maximum drift


demand expected for zone 3 soil type D for a stiff building.
Some walls, besides presenting concrete cover spalling and longitudinal bar buckling, showed overall buckling promoted by small
wall thickness. To emphasize this situation, the code requires a stability analysis for walls with thicknesses below lu/16, with lu
unsupported wall height.
Failure of splices in longitudinal reinforcement in walls was observed in a few cases. Considering the need to maintain the walls
integrity as a load transfer mechanism through the longitudinal
reinforcement, especially in the conned core of the critical
section, when the amount of longitudinal reinforcement exceeds
2.8/fy or presents an small coating (coverrebar 6 2db,long), a
minimum quantity of transversal reinforcement is required
(Atr,min = Abfyls/fytld, with Ab and fyl the longitudinal reinforcement
area and yield stress, s and fyt the transversal reinforcement spacing and yield stress, and ld the longitudinal reinforcement development length) to be able to transmit forces between the overlapped
bars. In this case, it is required that the force capacity by the transversal reinforcement in the overlapped area can at least balance
the component (parallel to the transversal reinforcement) of the
forces transmitted to the concrete (with inclined struts at 45)
due to the forces in the longitudinal bar in tension.
All this code requirements have been used by engineering companies during the past few months and they have reported that it
has resulted in a marginal increase of cost in RC shear wall building
construction.

tension to compression causing instability even before compressive strains are observed.
Where damage was produced by large compressive or tensile
strains that resulted in degrading response (softening post-yield
response) with localization of damage within a small area, a better
distribution of strain is expected for a non-degrading response, by
providing connement and/or buckling constraint. The better
strain distribution results in smaller maximum strain values,
reducing damage and making less probable the presence of fracture of longitudinal reinforcement due to low cycle fatigue.
Other limitations were also imposed to improve the concrete
placement and effectiveness of connement, as well as prevention
of global buckling. Several analytical and experimental programs
have been carried out lately to study, in more detail, aspects such
as overall response, effective connement, effect of discontinuities,
among others, by the author and other researchers, which would
allow for the checking and improvement of the code provisions.

Acknowledgements
This work is the summary of long discussions within the subcommittee for the RC design code committee formed under the
Construction Institute (Chile). The author of this publication, coordinator of the subcommittee, would like to thank the contribution
from several persons related to the academia and industry who
served as members and collaborated by enriching the discussion
with examples, case studies, and analysis of the new code requirements and its impact on Chilean construction. The members, in
alphabetical order, are: Marcial Baeza, Patricio Bonelli, Leopoldo
Breschi, Jorge Carvallo, Rodrigo Concha, Luis Daz, Augusto Holmberg, Ivn Hrepic, Matas Hube, Denis Jequier, Rodrigo Jordn, Marianne Kpfer, Mario Lafontaine, Ren Lagos, Alfonso Larran, Carl
Lders, Rodrigo Mujica, Miguel Sndor, Hernn Santa Mara, Eduardo Santos, Rodolfo Saragoni, Carlos Seplveda, and Rodrigo Vsquez. The author would also thank the wall specimen photos
provided by prof. John Wallace at UCLA.

9. Summary and conclusions

References

Modern RC building suffered severe damage in many cases, due


to axial load and bending in walls, after the Mw 8.8 earthquake in
2010 in Chile, expressed mainly as concrete cover spalling followed
by longitudinal boundary bar buckling and concrete crushing. The
absence of wall boundary detailing explains such behavior, but
older structures and also thousands of other buildings suffered
minor or no damage at all. The potential of the axial load to reduce
the displacement capacity is discussed, as well as other factors,
such as the impact of cross-section asymmetry.
The fundamental principles for the Chilean RC design code
changes are analyzed. Currently, a displacement-based approach
is used for detailing of the wall boundary elements, following the
ACI 318-08 requirements with minor changes. A new limit state
is established, providing a damage limitation. Compressive concrete strain cannot exceed 0.008, limiting indirectly the allowed
axial load, which was one of the potential precursors of the
damage.
Besides, restraining buckling is also provided by requiring
transversal reinforcement with a maximum spacing of 6db (db,
longitudinal bar diameter), in zones of potential yielding of
the longitudinal reinforcement, such that buckling due to tensioncompression cycles is minimized. This was required since,
in asymmetric cross-sections, it is possible to have relatively low
compressive strains such that no wall boundary detailing is
required, but buckling is expected upon reversal of cycles from

[1] Massone LM, Bonelli P, Lagos R, Lders C, Moehle J, Wallace JW. Seismic design
and construction practices for RC structural wall buildings. Earthq Spectra
2012;28(S1):S24556.
[2] NCh433.Of96. Earthquake resistant design of buildings. Instituto Nacional de
Normalizacin, INN 1996, 43pp [in Spanish].
[3] Caldern JA. Update on structural system characteristics used in RC buildings
in Chile. Civil Engineering Thesis, University of Chile; 2007. 76pp. [in Spanish],
[4] Wallace JW, Massone LM, Bonelli P, Dragovich J, Lagos R, Lders C, Moehle J.
Damage and implications for seismic design of RC structural wall buildings.
Earthq Spectra 2012;28(S1):S28199.
[5] Massone LM, Moroder D. Buckling modeling of reinforcing bars with
imperfections. Eng Struct 2009;31(3):75867.
[6] Lacaze C. Study and modeling of the impact of buckling in low cycle fatigue in
reinforcing bars for reinforced concrete. Civil Engineering Thesis, University of
Chile; 2009. 94pp [in Spanish].
[7] Brown J, Kunnath SK. Low-cycle fatigue failure of reinforcing steel bars. ACI
Mater J 2004;101(6):45766.
[8] Thomsen JH, Wallace JW. Displacement-based design of slender RC structural
walls-experimental verication. J Struct Eng, ASCE 2004;130(4):61830.
[9] Karamlou A, Kabir MZ. Experimental study of L-shaped slender R-ICF shear
walls under cyclic lateral loading. Eng Struct 2012;36:13446.
[10] Massone LM, Wallace JW. Load-deformation responses of slender reinforced
concrete walls. ACI Struct J 2004;101(1):10313.
[11] Cordero FA. Modeling of RC wall compression failure observed in the Mw 8.8
earthquake in 2010 in Chile. Civil Engineering Thesis, University of Chile;
2011. 105pp [in Spanish].
[12] Rodriguez ME, Botero JC, Villa J. Cyclic stressstrain behavior of reinforcing
steel including effect of buckling. J Struct Eng, ASCE 1999;125(6):60512.
[13] Moyer MJ, Kowalsky MJ. Inuence of tension strain on buckling of
reinforcement in concrete columns. ACI Struct J 2003;100(1):7585.
[14] Syntzirma DV, Pantazopoulou SJ, Aschheim M. Load-history effects on
deformation capacity of exural members limited by bar buckling. J Struct
Eng, ASCE 2010;136(1):111.

L.M. Massone / Engineering Structures 56 (2013) 13351345


[15] ACI 318-08: Building code requirements for structural concrete and
commentary. Committee 318. Farmington Hills, Michigan: American
Concrete Institute; 2008. 467pp.
[16] Wallace JW, Orakcal K. ACI 318-99 provisions for seismic design of structural
walls. ACI Struct J 2002;99(4):499508.
[17] Bohl A, Adebar P. Plastic hinge lengths in high-rise concrete shear walls. ACI
Struct J 2011;108(2):14857.
[18] Dazio A, Beyer K, Bachmann H. Quasi-static cyclic tests and plastic hinge
analysis of RC structural walls. Eng Struct 2009;31(7):155671.
[19] Paulay T, Priestley MJN. Seismic design of reinforced concrete and masonry
buildings. New York: John Wiley & Sons, Inc.; 1992. 768pp.
[20] Paulay T. The displacement capacity of reinforced concrete coupled walls. Eng
Struct 2002;24(9):116575.
[21] Zhang YF, Wang ZH. Seismic behavior of reinforced concrete shear walls
subjected to high axial loading. ACI Struct J 2000;97(5):73950.
[22] Su RKL, Wong SM. Seismic behaviour of slender reinforced concrete shear
walls under high axial load ratio. Eng Struct 2007;29(8):195765.
[23] Schlaich J, Schfer K, Jennewein M. Toward a consistent design of structural
concrete. PCI J 1987;32(3):75150.

1345

[24] GUC, seismological services at the University of Chile. <http://


www.sismologia.cl/>.
[25] Sawyer HA. Design of concrete frames for two failure stages. In: Proceedings of
the international symposium on exural mechanics of reinforced concrete,
ASCE-ACI, Miami, FL; 1964. p. 40531.
[26] NCh430.Of2008. Reinforced concrete-design and calculation requirements.
Instituto Nacional de Normalizacin, INN; 2008. 17pp [in Spanish].
[27] D.S. N 61 MINVU. Building seismic design code, replacing D.S N 117, 2010.
Chilean Ministry of Housing and Urbanism, Diario Ofcial; 13 December 2011
[in Spanish].
[28] D.S. N 60 MINVU. Reinforced concrete design code, replacing D.S N 118, 2010.
Chilean Ministry of Housing and Urbanism, Diario Ocial; 13 December 2011
[in Spanish].
[29] Estay CG. Characteristics of RC walls design in Chile. Civil Engineering Thesis,
University of Chile; 2008. 57pp [in Spanish].
[30] Saatcioglu M, Razvi SR. Strength and ductility of conned concrete. J Struct
Eng-ASCE 1992;118(6):1590607.

You might also like