Uploaded by Preston Varghese

Seismic-ResistantPrecastConcreteStructuresStateoftheArt

advertisement
See discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/322771671
Seismic-Resistant Precast Concrete Structures: State of the Art
Article in Journal of Structural Engineering · April 2018
DOI: 10.1061/(ASCE)ST.1943-541X.0001972
CITATIONS
READS
268
1,417
8 authors, including:
Yahya C. Kurama
Sri Sritharan
University of Notre Dame
Iowa State University
137 PUBLICATIONS 4,037 CITATIONS
241 PUBLICATIONS 5,133 CITATIONS
SEE PROFILE
SEE PROFILE
Robert Fleischman
Jose Ignacio Restrepo
The University of Arizona
University of California, San Diego
63 PUBLICATIONS 1,272 CITATIONS
216 PUBLICATIONS 5,573 CITATIONS
SEE PROFILE
All content following this page was uploaded by Ned Cleland on 01 May 2023.
The user has requested enhancement of the downloaded file.
SEE PROFILE
60th Anniversary State-of-the-Art Reviews
Seismic-Resistant Precast Concrete Structures:
State of the Art
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
Yahya C. Kurama, M.ASCE 1; Sri Sritharan, M.ASCE 2; Robert B. Fleischman, M.ASCE 3;
Jose I. Restrepo, M.ASCE 4; Richard S. Henry 5; Ned M. Cleland, M.ASCE 6;
S. K. Ghosh, F.ASCE 7; and Patricio Bonelli 8
Abstract: Precast concrete facilitates a construction method using durable and rapidly erectable prefabricated members to create costeffective and high-quality structures. In this method, the connections between the precast members as well as between the members
and the foundation require special attention to ensure good seismic performance. Extensive research conducted since the 1980s has led
to new precast concrete structural systems, designs, details, and techniques that are particularly suited for use in regions of high seismic
hazard. This paper reviews the state of the art of these advances, including code developments and practical applications, related to four
different systems: (1) moment frames; (2) structural walls; (3) floor diaphragms; and (4) bridges. It is concluded from this review that the
widespread use of precast concrete in seismic regions is feasible today and that the jointed connection innovation introduced through precast
research leads to improved seismic performance of building and bridge structures. DOI: 10.1061/(ASCE)ST.1943-541X.0001972. © 2018
American Society of Civil Engineers.
Author keywords: Code developments; Earthquake resistance; Emulated; Joints; Posttensioning; Precast prestressed concrete;
Prestressing; Pretensioning; Seismic design; Unbonded.
Introduction
In precast concrete construction, structural members (e.g., beams,
columns, wall panels, and floor units) and architectural members
(e.g., cladding) are produced in a manufacturing facility, transported to the structure site, and erected and connected in place.
The cost effectiveness and high quality of this construction method
have resulted in its widespread use, including a number of countries
with high seismicity (e.g., New Zealand, Japan, and Chile).
The seismic behavior of precast building and bridge structures is
highly dependent on the characteristics (i.e., strength, stiffness, and
deformation capacity) of the connections between the precast structural members and between the members at the base and the foundation. Poor performance of precast buildings in past earthquakes
1
Professor, Dept. of Civil and Environmental Engineering and
Earth Sciences, Univ. of Notre Dame, 156 Fitzpatrick Hall, Notre Dame,
IN 46556 (corresponding author). E-mail: ykurama@nd.edu
2
Wilkinson Chair Professor in Engineering, Dept. of Civil, Construction
and Environmental Engineering, Iowa State Univ., Ames, IA 50011.
3
Professor, Dept. of Civil Engineering and Engineering Mechanics,
Univ. of Arizona, Tucson, AZ 85721.
4
Professor, Dept. of Structural Engineering, Univ. of California at
San Diego, La Jolla, CA 92093.
5
Senior Lecturer, Dept. of Civil and Environmental Engineering, Univ.
of Auckland, Auckland 1010, New Zealand.
6
President, Blue Ridge Design, Inc., 3 W. Piccadilly St., Winchester,
VA 22601.
7
President, S. K. Ghosh Associates, Inc., 334 East Colfax St., Unit E,
Palatine, IL 60067.
8
Professor, Dept. of Civil Engineering, Universidad Técnica Federico,
Santa María, Valparaiso, Chile.
Note. This manuscript was submitted on January 21, 2017; approved on
August 22, 2017; published online on January 17, 2018. Discussion period
open until June 17, 2018; separate discussions must be submitted for individual papers. This paper is part of the Journal of Structural Engineering, © ASCE, ISSN 0733-9445.
© ASCE
has often been attributed to poorly designed and/or poorly built
connections (fib 2003). Accordingly, the need to develop new
seismic building code requirements for precast construction that
specifically address the unique role of the precast connections
was recognized in the United States and elsewhere beginning in the
1980s (e.g., Englekirk 1986; Hawkins and Englekirk 1987; Park
1995).
In the United States, the first major research study with a focus
on the seismic performance of precast building structures was
undertaken during the 1990s at the National Institute of Standards
and Technology (e.g., Cheok and Lew 1993). This study focused
specifically on building moment frames. A major U.S.–Japan cooperative research program on PREcast Seismic Structural Systems
(PRESSS) was initiated in the early 1990s (Priestley 1991). This
program comprised many individual research projects, followed
by the design and testing of a culminating five-story precast building structure (Nakaki et al. 1999). These efforts formed the basis of
a number of other subsequent research studies in the United States.
In the 2000s, a large research program was conducted on precast
floor diaphragms (Fleischman et al. 2013). In comparison with
buildings, developments in seismic precast concrete bridges were
pursued less aggressively until advancements occurred in accelerated bridge construction in nonseismic regions.
Extensive research and development on seismic precast building
structures have also been undertaken outside the United States
(again, research in bridge structures has been more recent). Among
these, significant research in Japan (since the U.S.–Japan PRESSS
program) includes the testing of a full-scale four-story building
using precast posttensioned concrete members at the E-Defense
shaking table facility (Nagae et al. 2014). In New Zealand, research
programs in the 1980s and 1990s led to the development of guidelines for structural precast concrete (CAE 1999) and the implementation of precast construction into building design standards
(Park 1995). In addition, research into floor diaphragms resulted
in changes in diaphragm seismic design and connection detailing
03118001-1
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
gap
lateral load
wall
foundation
lateral displacement
(a)
extensive
cracking
wall
lateral load
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
(Fenwick et al. 2010). In Europe, a collaborative five-year research
project called SAFECAST was conducted to study the behavior of
precast concrete buildings under earthquake loading (e.g., Negro
et al. 2013; Psycharis and Mouzakis 2012). China, which suffered
significant precast building failures in the 1976 Tangshan earthquake (Housner and Lili 2002) and subsequently moved away from
precast, has recently been making efforts to revive this construction
form (Seeber 2014).
The previous research on seismic precast building and bridge
structures, spanning more than three decades, has in turn led to
changes to the relevant governing codes and successful implementations in regions of high seismicity. This paper, which provides a
review of these advances, is organized into the following major sections on seismic-resistant precast concrete: (1) general concepts;
(2) development of seismic provisions in design codes; (3) connections, splices, and anchorages; (4) moment frames; (5) structural
walls; (6) floor diaphragms; and (7) bridges. Although this review
paper attempts to include all major relevant research advances in
the United States and abroad, the discussion of code developments
is focused mostly on the United States. Furthermore, on-site precast
(e.g., tilt-up) and architectural precast are outside the scope of the
paper.
foundation
lateral displacement
(b)
Fig. 1. Unique behavior of unbonded PT structures: (a) unbonded PT
wall with gap opening; (b) conventional monolithic cast-in-place
reinforced concrete wall
General Concepts
Precast concrete construction for seismic systems can be broadly
classified into two types, emulative and jointed. Emulative construction uses connections that are designed and detailed to make
the performance (in terms of lateral strength, stiffness, and energy
dissipation) of the precast structure comparable to that of an equivalent, conventionally designed, and properly detailed cast-in-place
monolithic reinforced concrete structure (Ericson and Warnes
1990). Emulative connections are further divided into two types:
ductile and strong. Structures with ductile connections are designed
to undergo flexural yielding and form ductile plastic hinges in the
connections across precast member-to-member or member-tofoundation joints, whereas structures with strong connections are
designed to experience flexural yielding within the precast members at preselected and appropriately detailed locations adjacent to
or away from the joints. This paper refers to the region at the intersection of (i.e., between) precast members as a joint, and refers to
the assembly of hardware and anchors across a joint as a connection. Capacity-design principles are used to ensure that strong connections remain essentially in the linear-elastic range of behavior,
while plastic hinges fully develop elsewhere in the structure.
Jointed construction (also referred to as nonemulative detailing
in the literature) uses precast connection concepts that are distinctly
different from emulative connections. In this approach, the nonlinear rotations of the structure are deliberately concentrated at the
ends of the precast members in the joint regions (through controlled
rocking at the joint interface), without causing significant inelastic
behavior (i.e., damage) in the members. This unique behavior
has been achieved by using unbonded posttensioning (PT) steel
(often using multistrand tendons) as the primary reinforcement
in moment frames and walls, as demonstrated in the PRESSS program (e.g., Nakaki et al. 1999; Priestley et al. 1999) and subsequent
research.
The objectives of jointed connections, in not attempting to emulate the lateral load behavior of monolithic cast-in-place structures, are twofold. First, the concentrated rotations (also referred
to as rocking rotations) at the jointed connections significantly reduce (or eliminate) tension damage in the precast members by
allowing the structure to undergo large lateral displacements with
© ASCE
little tensile stresses in the concrete. Second, the PT force provides
a large restoring effect that facilitates recentering of the structure at
the end of the seismic loading (often referred to as self-centering
capability).
Fig. 1 illustrates how an unbonded PT connection concentrates
rotations through the opening of a discrete gap at the joint (which is
controlled by the tension resistance from PT) by comparing the
behavior of a typical unbonded PT wall [Fig. 1(a)] with the development of extensive cracks (and associated yielding of reinforcement) in a conventionally reinforced concrete wall [Fig. 1(b)]. The
changes in the neutral axis depth through this gap opening and, to
some extent, the development of nonlinear-inelastic compressive
behavior of concrete in the extreme compression corners (toes)
of the member, allow the moment-rotation behavior of the connection to be nonlinear without causing significant inelastic behavior
(i.e., damage) or strength degradation in the precast member. The
gap opening initiates when the moment at the joint overcomes the
precompression from PT, and for vertical members (e.g., columns,
walls, and piers), any gravity loads. The deliberate lack of bond
between the PT steel and concrete along the precast members significantly reduces the strain build-up in the tendon (by uniformly
distributing the steel strains over the length of the tendon). This
allows the PT steel to remain essentially linear-elastic until the precast member/structure reaches a target lateral displacement, thereby
maintaining the initial prestress force when the structure returns to
zero displacement after being subjected to large nonlinear displacements. Upon unloading, the force in the PT tendons (together with
any gravity loads if they are vertical members) facilitates closing of
this gap. When the loading is reversed, the gap opens in the other
direction. The width of the gap is controlled by the tension resistance provided by the PT steel, causing controlled rocking at the
joint.
When the jointed connection is established solely using unbonded PT steel to provide tensile resistance, it leads to nearly
nonlinear-elastic behavior under reversed-cyclic loading [Fig. 1(a)],
with little residual lateral displacements and limited hysteretic energy dissipation. As a major disadvantage, the significantly smaller
energy dissipation may result in increased seismic displacement
03118001-2
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
demands for the precast structure (Priestley and Tao 1993; Farrow
and Kurama 2003; Seo and Sause 2005). Thus, to comply with seismic design requirements, precast structures with jointed connections are often designed with external or internal supplemental
energy dissipation components (e.g., through yielding or friction/
sliding), while maintaining their unique features, as described for
different structural systems subsequently in this paper. Another potential drawback of jointed connections is that the compression toe
of the rocking precast member may experience damage through
crushing of the concrete. Typically, this is limited to the cover concrete as long as the core region is sufficiently confined (e.g., Smith
et al. 2013, 2015). The cover concrete damage can also be prevented
or significantly reduced (e.g., Nazari et al. 2017) by using engineered cementitious composites (ECCs) in the vicinity of the joint,
steel or fiber-reinforced polymer (FRP) jacketing, steel armoring,
and, more recently, by using rubber sheets, polymer joints, and
rubber bearing joints at the appropriate locations.
Development of Seismic Provisions
in Design Codes
Seismic design of precast concrete structures in most U.S. jurisdictions today is in accordance with the International Building
Code IBC-15 (IBC 2014), which adopts two key standards, ASCE/
SEI 7-10 (ASCE 2013) and ACI 318-14 (ACI 2014). Other important resource documents are the National Earthquake Hazards
Reduction Program (NEHRP) Provisions (BSSC 2015), PCI
Design Handbook (PCI 2010), and PCI Seismic Design Manual
(Cleland and Ghosh 2012).
Seismic Force Resisting Systems for Building
Structures
The ACI 318 code included seismic design provisions for precast
concrete for the first time in 2002, largely based on the 2000
NEHRP Provisions (BSSC 2001). These provisions provided emulative design specifications applicable to special moment frames
and structural walls. Design of moment frames involving jointed
connections was allowed based on a validation testing document,
the current designation of which is ACI 374.1-05 (ACI 2005). ACI
318-08 (ACI 2008) added jointed design provisions for special
structural walls, based on another validation testing document,
ACI ITG 5.1-07 (ACI 2007). Hawkins and Ghosh (2006), Ghosh
(2004), and D’Arcy et al. (2003) provided more information on the
historical evolution of precast seismic design provisions for buildings in the United States. Specific code provisions for different seismic building structural systems appear in the sections “Moment
Frames” and “Structural Walls.”
Among other countries, the New Zealand Concrete Structures
Standard, NZS 3101 (Standards New Zealand 2006), contains specific provisions for the seismic design of precast concrete members,
including both emulative and jointed construction. Appendix B of
NZS 3101 includes special provisions for the seismic design of
ductile jointed precast structural systems, which were first introduced in the 2006 edition. In addition, there are guidelines specific
to precast concrete structures, such as CAE (1999) and the PRESSS
design handbook (Pampanin et al. 2010).
Floor Diaphragms for Building Structures
With regard to precast floor diaphragms in the United States, the
ASCE/SEI 7-16 (ASCE 2016) standard adopted two new sections
(i.e., Sections 12.10.3 and 14.2.4) based on significant updates to
diaphragm design in the 2015 NEHRP provisions (BSSC 2015).
© ASCE
ASCE/SEI 7-16 Section 12.10.3 presents an alternative determination of the diaphragm design force based on rational methods,
whereas Section 14.2.4 contains options for the seismic design of
diaphragms based on Zhang and Fleischman (2016). Because of
the recent nature of these developments, the “Floor Diaphragms”
section of this paper includes a detailed treatment of this topic.
Similar new recommendations have been developed in New Zealand
[NZS 1170.5 (Standards New Zealand 2004) and NZS 3101
(Standards New Zealand 2006)].
Bridge Structures
Code provisions for using precast concrete in seismic bridge design
have not fully evolved. This lack of code provisions stems from the
infrequent use of precast concrete historically in seismic design of
bridges. Specifically, the use of precast girders in bridges has been
hampered by two design criteria: (1) assuming that a positive moment connection between precast girders and the bent cap beam
cannot be reliably developed, formation of a plastic hinge at the
column (or pier) top is not permitted, making this construction option less cost effective compared with the cast-in-place alternative
(Caltrans 2013; Vander Werff et al. 2015); and (2) a connection
involving mild steel reinforcement is required between the cap
beam and girders to ensure satisfactory shear transfer across the
cap beam-to-girder joint when vertical acceleration exceeds 0.25 g
(Caltrans 2013). As such, the use of precast girders in bridge structures, when it occurs, is often motivated by overcoming construction challenges associated with cast-in-place concrete.
Research to address the aforementioned concerns and the use of
other forms of precast concrete in seismic bridge design have been
completed in recent years. Outcomes of these studies and experiences using precast concrete in selected prototype bridges have
led to the development of design guidelines by state agencies
(e.g., Departments of Transportation in California, Utah, and
Washington). Amalgamation of these guidelines can eventually
lead to generalized seismic design provisions.
Connections, Splices, and Anchorages
Connections in seismic precast structural systems are typically
completed in the field during erection. A construction space is often
created at the joints for tolerances and alignment purposes. This
space is filled with high-strength, nonshrink grout, sometimes including synthetic fibers. The connections across the joints can be
made using an assortment of hardware embedded in the precast
members, anchorage of deformed mild steel reinforcement, and/or
bonded or unbonded PT (PCI 2010; PTI 2006). ACI 550R-96 (ACI
2001) gives connection details specific to emulative construction.
Hardware connections and bonded PT are typically used in emulative construction. Hardware connections, which can be bolted or
welded, include various types of embedded plates and commercial
inserts. For many types of hardware connections, completing the
load path depends on anchorage to concrete, which is not discussed
herein but is provided in Chapter 17 of ACI 318-14.
Anchorage of deformed mild steel reinforcing bars is important
in both emulative and jointed precast construction. Anchorage of
these bars can be achieved using high-strength, nonshrink grout
(i.e., a prepackaged mixture of cement, sand, water, and admixtures
for bonding reinforcement in sleeves or ducts) or in-situ concrete, or
mechanical splicing in the field. Normally, the lap lengths required
in accordance with Section 25.5 of ACI 318-14 are too long for the
assembly of precast members. Thus, mechanical splice devices with
short projecting lengths of bars are often used across both horizontal and vertical joints. These splice devices are typically rigid;
03118001-3
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
hence, unless detailed carefully, they can force the bar elongation to
occur over a very short length just outside the splice, which can lead
to high local strains and early bar fracture (Marsh et al. 2011). This
can be mitigated by the use of debonded lengths (or stretch lengths)
to distribute the steel strains over a finite length of the bars in jointed
precast construction, as has been demonstrated for both frame and
wall systems (e.g., Priestley et al. 1999; Smith et al. 2011).
Type 2 splices are commonly used for deformed mild steel
reinforcing bars in precast concrete construction. According to
ACI 318-14 Section 18.2.7.1, Type 2 splices for seismic design
must be “capable of developing the specified tensile strength of the
spliced bars.” ACI 318-14 permits these splices to be used at any
location, including yielding regions, except for ductile connections
in special moment frames, where they are not permitted closer than
half the beam depth from the column face. According to Section
R18.2.7.1 of ACI 318-14, the requirements for Type 2 splices are
intended to result in a device that is “capable of sustaining inelastic
strains through multiple cycles.” However, grouted Type 2 splices
have experienced failure under cyclic loading due to bond pullout
in jointed precast walls (Smith and Kurama 2014) tested in accordance with ACI ITG 5.1-07 (ACI 2007). Research is currently
underway (Aragon et al. 2017) to validate an improved grouted
connector for ductile reinforcing bars to not only meet Type 2 requirements, but also sustain extreme postyield cyclic bar strains
required in jointed precast seismic systems.
Unbonded PT is common in jointed construction. Anchorages
for unbonded PT strands have been the subject for several studies
(Walsh and Kurama 2010, 2012; Sideris et al. 2014a; Musselman
et al. 2015). These studies have shown that premature wire fractures
of unbonded seven-wire strand can occur inside industry-standard
PT anchorages. To prevent this behavior, the studies have recommended maximum allowable strand strains for use in the design of
jointed precast structures. Walsh et al. (2015) showed that higherperforming PT anchors that allow greater strand strains without
fracture can be achieved by slightly changing the anchor wedge
geometry.
Moment Frames
This section summarizes the developments associated with using
precast concrete in seismic design of moment frames.
Moment Frames with Emulative Connections
Sections 18.9.2.1 and 18.9.2.2 of ACI 318-14 provide emulative
design provisions for “special” precast concrete moment frames.
As with monolithic cast-in-place frames, column bases and beam
ends are typically selected as the plastic hinge locations in emulative precast frames. Providing adequate confinement reinforcement
in these critical regions ensures formation of ductile plastic hinges.
Nonlinear action in other regions is prevented by using capacitydesign principles. Designers must carefully select locations of
strong connections or take other measures, such as debonding of
reinforcing bars in highly stressed regions, to avoid strain concentrations that have the potential to cause premature fracture of
reinforcement.
Development of different emulative connection concepts
(e.g., Park and Bull 1986; Park 1990; Restrepo et al. 1995) and
related design documents (e.g., AIJ 2000; Park 1995; ACI
2001) have led to the frequent use of these systems in Japan,
New Zealand, and to some extent, in the United States. Frame structures may use cast-in-place or precast columns with precast or partially precast beams. Partially precast beams may use exposed
reinforcement or a precast shell for the beams. Figs. 2(a–c) show
© ASCE
examples of emulative beam-to-column connections. Typical emulative frame construction built from the 1980s onward generally
performed well during the Canterbury earthquakes in New Zealand
(Fleischman et al. 2014). Walsh et al. (2016) tested damaged precast
frames extracted from a 20-story building to assess their residual
capacity and reparability. These tests confirmed the behavior of emulative precast frames tested previously in the laboratory (Restrepo
et al. 1995) and also showed that flexural hinging in the beams
can be repaired with no significant reduction in seismic capacity.
In Japan, the emulative concept has generally favored the use of
bonded PT (Nishiyama 1990). This framing system has used multistory columns and single-bay beams with bonded PT establishing
the beam-to-column connections [Fig. 2(d)]. Unlike emulative
connections with deformed mild steel reinforcement, this framing
concept provides reduced energy dissipation because of the use of
PT steel. Furthermore, inelastic strains in the PT steel can cause
prestress losses under cyclic loading. The plastic hinging in this
system will occur at the beam ends and likely penetrate into the
joint (fib 2003). As a result, debonding of the strands within the
plastic hinge region is possible, minimizing the loss of prestressing.
Muguruma et al. (1995) reported that more than 150 moment
frames designed with this concept performed well in the 1995 Kobe
earthquake in Japan. Importantly, these frames recentered at the end
of ground shaking, enabling buildings to be functional following
the Kobe earthquake, despite the potential for prestress loss and
low amount of energy dissipation.
Moment Frames with Jointed Connections
According to Section 18.9.2.3 of ACI 318-14, special precast
moment-resisting frames with jointed connections not satisfying
Sections 18.9.2.1 and 18.9.2.2 of ACI 318-14 are permitted, provided they satisfy ACI 374.1-05 (ACI 2005). Section 18.9.2.3 imposes additional criteria related to the use of representative details
and materials if test validation is required, and provides a guiding
design procedure to identify the load path or mechanism by which
the frame resists gravity and earthquake effects.
Moment frames with jointed connections frequently use multistory columns and single-bay beams, although use of single-story
columns and multibay beams has also been suggested. Fully or
partially unbonded PT through the beams is the primary feature in
these frames, which are typically designed to remain linear-elastic
under the design-level earthquake. When the frames are constructed
with single-bay beams, unbonded PT is used to establish the connections between the columns and beams. Because the PT steel is
designed to remain linear-elastic, a jointed frame provides lower
energy dissipation capacity than frames with bonded prestressing
(El-Sheikh et al. 1999). Consequently, supplemental damping for
frames with unbonded PT has been tested (e.g., Nakaki et al. 1999;
Priestley et al. 1999; Morgen and Kurama 2004, 2007, 2008).
The hybrid frame concept is the most successful jointed frame to
date. The term hybrid refers to the combined use of unbonded PT
steel and deformed mild steel reinforcing bars to form the beam-tocolumn connections [Fig. 3(a)]. The PT steel is placed at or near the
midheight of the beam, whereas the mild steel bars are placed closer
to the top and bottom of the beam. Under lateral loads, gaps form at
the joints at the beam ends when the moment at the connection exceeds the decompression moment. The mild steel bars contribute to
the moment resistance and are designed to yield and provide energy
dissipation during the gap opening–closing behavior (i.e., rocking
rotations) under cyclic loading. To control the maximum tension
strain demands on the mild steel bars and prevent their premature
low-cycle fatigue fracture, these bars are debonded over a short
length (stretch length) in or adjacent to the connection region.
03118001-4
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
Fig. 2. Example emulative and bonded PT beam-to-column connections: (a) emulative connection with in-situ concrete; (b) emulative connection
with grouted bars; (c) emulative corner beam-to-column connection; (d) bonded PT beam-to-column connection [(a, b, and d) adapted from fib
Bulletin 27: “Seismic design of precast concrete building structures“ (October 2003), with permission from the International Federation for Structural
Concrete (fib); (c) image by Jose I. Restrepo]
Rebar debonded
over a short length
Grout with fiber
Mild steel
reinforcement
Unbonded posttensioning
(b)
(a)
Fig. 3. Example of a jointed frame beam-to-column connection and its use in a system: (a) jointed hybrid beam-to-column connection
(image courtesy of S. Nakaki, Nakaki Structural Design, Tustin, CA); (b) three-story hybrid frame tested in the PRESSS building (image by
Sri Sritharan)
As part of the concept development, several interior hybrid moment frames were tested at the National Institute for Standards and
Technology (Cheok and Lew 1993; Stone et al. 1995). Although
these tests used partially bonded PT strands, the strands were unbonded in the critical region that included the column and a short
distance within the beam on either side of the column. In the culmination of the PRESSS program, a five-story moment frame
© ASCE
consisting of several different jointed connections was subjected
to pseudodynamic testing, which included a three-story, two-bay
hybrid frame [Fig. 3(b)] (Priestley et al. 1999; Sritharan 2002).
These tests validated the superior performance of this jointed system to provide (1) sufficient energy dissipation, (2) minimal structural damage, and (3) recentering capability. A design document,
ACI ITG 1.2-03 (ACI 2003), was subsequently developed for
03118001-5
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
grew in seismic regions, demonstrating the practical viability for
widespread use in low-rise and mid-rise buildings. Shutt (1997)
explained the advantages of the hybrid frame, particularly for
parking structures, citing several applications [e.g., Fig. 4(a)].
The hybrid frame concept was also used in the 39-story Paramount
apartment building [Fig. 4(b)] in San Francisco (Englekirk 2002)—
the tallest concrete-framed structure in a high-seismic zone in the
United States at the time of completion in 2002. The next section,
“Structural Walls,” provides two other applications of hybrid
frames.
Fig. 4. Unbonded PT frames: (a) parking structure under construction
at University of California, Davis Medical Center (image courtesy of
Clark Pacific, West Sacramento, CA); (b) 39-story Paramount Tower in
San Francisco (image courtesy of J. Sanders, Charles Pankow Builders,
Pasadena, CA)
Structural Walls
Structural walls are a commonly used lateral load resisting member
in building design due to their large lateral stiffness, strength, and
cost effectiveness. Precast walls can be constructed using rectangular uniformly thick wall panels with emulative or jointed connections. Possible configurations include single or coupled walls, with
each wall comprising individual or multiple panels (Fig. 5). In a
multipanel wall, vertical stacks of panels (typically one story high)
are connected across horizontal joints. Coupled walls consist of two
or more vertically oriented walls connected together in the same
plane by connectors or coupling beams.
hybrid moment frames with jointed precast beam-to-column connections. As required by Section 18.2.1.7 of ACI 318-14, the ACI
ITG 1.2-03 guidelines ensure that hybrid frames have strength and
toughness at least equal to those of comparable monolithic frames
satisfying Chapter 18 of ACI 318-14.
One issue with hybrid moment frames is that the gap opening
at the beam-to-column joints can damage the adjacent floor slab.
This has led to configurations that reduce damage to the floor slab
during cyclic loading, such as the TYC-gap connection used in the
PRESSS test building (Priestley et al. 1999; Sritharan 2002) and a
slotted beam detail tested by Muir et al. (2012).
Since the development of design guidelines for hybrid frames,
precast frame research for seismic applications has subsided. At the
same time, hybrid frame applications in commercial structures
Structural Walls with Emulative Connections
ACI 318-14 Section 18.11.2.1 requires special emulative precast
walls to satisfy the requirements in Section 18.10 for special monolithic cast-in-place walls. It is also required that in the connections
between the wall panels, or between the base panel and the foundation, yielding be limited to steel components or reinforcement,
and that the parts of the connection that are not intended to yield
unbonded
PT tendon
PT anchorages
unbonded
PT tendon
wall panel
horizontal
joint
wall panel
wire mesh
wire mesh
fiber-reinforced
grout
bonded mild
steel bar
plastic sleeve
confined
concrete
bonded mild
steel bar
plastic
sleeve
fiber-reinforced
grout
confined
concrete
plastic
sleeve
base panel
foundation
(a)
(b)
coupling
beam
coupling
connector
unbonded
PT tendon
(c)
(d)
Fig. 5. Precast walls: (a) single panel wall; (b) multipanel wall; (c) coupled wall with vertical joint; (d) coupled wall with coupling beams
© ASCE
03118001-6
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
be designed for at least 1.5 times the strength of the yielding parts
of the connection. To further ensure postyield ductility capacity,
IBC-15 (IBC 2014) Section 18.5.2.2 adds “connections that are designed to yield shall be capable of maintaining 80 percent of their
design strength at the deformation induced by the design displacement or shall use Type 2 mechanical splices.”
Early seismic research on emulative walls was conducted on systems with platform-type horizontal strong connections. In this system
(Martin and Korkosz 1982; ACI 2001), horizontal hollow-core slab
panels at the floor and roof levels are supported on bearing pads on the
wall panel below, and the upper wall panels rest on the platform
formed by the hollow-core panels, with the resulting connections
made using deformed reinforcing bars, bonded prestressing steel,
or mechanical connectors. Oliva et al. (1989, 1990) and Armouti
(1993) conducted large-scale experiments on precast walls typical
of platform-type construction, which showed that gap opening
(i.e., rocking rotations) along the horizontal joints is more desirable
as the primary mode of lateral deformation than is shear slip at the
joints. The specimens with bonded PT connections experienced debonding of the PT steel, which demonstrated beneficial effects of this
unintentional loss of bond in limiting the yielding of the PT steel and
reducing the flexural cracking of the wall panels. These observations
are consistent with the aforementioned benefits of jointed connections, in which the PT steel is intentionally unbonded, and formed
the basis for subsequent research on jointed unbonded PT walls.
Emulative walls can also be constructed by stacking precast panels on top of each other, without the floor slab in between, resulting
in simpler horizontal connections than in the platform-type system.
These connections between the wall panels and also to the foundation are typically achieved through the use of grouted corrugated
metal ducts or proprietary grout sleeves to splice deformed mild
steel reinforcing bars (Fig. 6) (ACI 2001; CAE 1999; Park 1995).
Concrete confinement reinforcement similar to that used in cast-inplace monolithic wall construction (or other special detailing) is
placed at the wall toes to ensure that the large compressive strain
demands can be sustained in this region under seismic loading.
A recent review of constructed precast walls in New Zealand
showed that grouted connections are common for both panel-topanel and panel-to-foundation connections, with the connection
flexural capacity often less than that of the panel (Seifi et al. 2016).
Seifi et al. (2017) investigated the robustness of grouted connections in precast panels with a single layer of reinforcement in a
series of tests, and observed reduced connection damage when
the metal ducts containing the connection reinforcement were confined with transverse reinforcement.
Fig. 6. Typical grouted wall panel-to-panel connections (adapted from
CAE 1999)
© ASCE
Structural Walls with Jointed Connections
Crisafulli et al. (2002) and Smith et al. (2013) tested jointed precast
walls (also referred to as rocking walls in the literature) featuring
gap opening connections with grouted mild steel reinforcing bars
across the wall-to-foundation joint. By allowing gap opening across
the wall-to-foundation joint, the amount of reinforcement in the
precast wall is significantly reduced as compared with monolithic
walls designed for seismic resistance. However, the tests by Smith
et al. (2013) showed that these walls can suffer from excessive uplift, shear slip, and lateral strength and stiffness degradation due to
the accumulation of plastic (i.e., residual) tensile strains in the connection reinforcement under reversed-cyclic loading.
The use of uncoupled precast walls featuring gap opening
across horizontal connections using only unbonded PT steel has
been studied by many researchers (Kurama et al. 1999a, b, 2002;
Galusha 1999; Allen and Kurama 2002a, b; Holden et al. 2003;
Perez et al. 2007, 2013; Erkmen and Schultz 2009; Henry et al.
2012; Belleri et al. 2014b). Because of the low energy-dissipation
capacity associated with unbonded PT connections, a number of
researchers have investigated the use of supplemental energy dissipation components in uncoupled precast walls (Kurama 2000,
2001; Ajrab et al. 2004; Restrepo and Rahman 2007; Marriott et al.
2008). Alternatively, Nazari et al. (2017) suggested that the design
base shear force can be adjusted to produce satisfactory results
for these systems despite the low hysteretic energy dissipation.
In coupled walls, vertically jointed walls with energy-dissipating
coupling details (Priestley et al. 1999; Perez et al. 2004a, b; Aaleti
and Sritharan 2009; Rahman and Sritharan 2015) as well as a column-wall-column system known as precast wall with end columns
(PreWEC) (Sritharan et al. 2015; Twigden et al. 2017) have been
studied [Figs. 7(a and b), respectively]. These systems use ductile
vertical joint connections to intentionally yield or slip during wall
rocking and promote the easy replacement of the energy dissipating
components if needed after an earthquake. Research has also investigated the behavior and design of unbonded PT precast coupling
beams for coupled walls [Fig. 5(d)] (Weldon and Kurama 2007,
2010, 2012).
ACI 318-14 Section 18.11.2.2 provisions permit the jointed design of special precast walls using unbonded PT and satisfying the
requirements of ACI ITG 5.1-07 (ACI 2007). ACI ITG 5.1 is a test
validation document, which was developed based on a provisional
standard and commentary produced by Hawkins and Ghosh
(2004), similar to the path that led to ACI 374.1-05 (ACI 2005)
for jointed special precast moment frames. ACI ITG 5.1 focuses
on uncoupled hybrid walls (with energy dissipating mild steel deformed reinforcement at the wall-to-foundation connection, similar
to the hybrid frame concept described previously) as well as vertically coupled walls (with energy dissipating coupling details, for
example, as used in the PRESSS program; Priestley et al. 1999;
Nakaki et al. 1999). The commentary to ACI 318-14 Section
18.11.2.2 refers to ACI ITG 5.2-09 (ACI 2009), which is a document for the design of special unbonded PT walls (both uncoupled
hybrid walls and coupled walls with energy dissipating coupling
details), comparable to ACI ITG 1.2-03 (ACI 2003) for the design
of special hybrid moment frames.
The hybrid wall system using unbonded PT together with energy dissipating deformed mild steel reinforcement across horizontal joints [Figs. 5 (a and b)] has been investigated extensively
(Rahman and Restrepo 2000; Holden et al. 2001, 2003; Kurama
2002, 2005; Restrepo 2003; Smith et al. 2011, 2013, 2015; Smith
and Kurama 2014). Smith and Kurama (2014) provided design and
detailing recommendations for hybrid walls (in addition to those
available in ACI ITG 5.2-09) based on testing of walls per ACI
03118001-7
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
Fig. 7. Precast walls: (a) PRESSS coupled wall (image by Sri Sritharan); (b) PreWEC system; (c) hybrid wall (reprinted from Smith et al. 2015,
with permission from the American Concrete Institute)
ITG 5.1-07 (ACI 2007). These recommendations ensure that hybrid
walls satisfy ACI 318-14 and ACI ITG 5.1-07 requirements for special RC structural walls. In multipanel walls with horizontal joints,
in addition to the design of the critical base-panel-to-foundation
connection, the upper panel-to-panel connections also must be designed adequately. Smith and Kurama (2014) described a methodology to prevent nonlinear behavior at the upper joints of hybrid
walls, thereby concentrating all gap opening at the base joint.
Applications of unbonded PT wall systems have been realized in
the Dominican Republic and New Zealand. The New Zealand applications include buildings throughout the country that have used
combinations of unbonded PT walls and frames. These systems
have included single unbonded PT wall panels as well as vertically
jointed walls and walls with steel coupling beams. A four-story
medical building in Christchurch that used unbonded PT walls
and hybrid frames [Fig. 8(a)] survived the 2010–2011 Canterbury
earthquakes without any significant damage (Pampanin et al.
2011). A five-story building with a similar configuration was under
construction during the 2010 Chile Earthquake [Fig. 8(b)] (Ghosh
and Cleland 2012). The erection of the lateral-load resisting system
for this building was complete, but the building was unfinished
at the time of the earthquake. Sritharan et al. (2015) argued that
the vertically jointed walls used in these buildings are not as
cost efficient as cast-in-place reinforced concrete walls. The
PreWEC system or uncoupled hybrid walls may be more efficient
because the wall panel length (and moment arm) is maximized
© ASCE
rather than dividing the wall length into shorter multiple wall panels
with reduced flexural capacities.
Floor Diaphragms
Floor (and roof) systems play a key role by providing diaphragm
action, thus uniting and transferring seismic forces to the lateral
load-resisting members (walls and moment frames). A key consideration for diaphragms in precast structures is the force transfer
across the joints between the precast floor units.
Precast floor units are typically double tees for long spans,
hollow core for medium spans, and can be ribs or flat slab sections
for shorter spans. Floor systems in precast structures can be topped
(i.e., provided with a cast-in-place topping over the precast units) or
untopped (in which the precast floor units are often provided with a
thicker flange region). Precast floor construction in terms of cross
sections and detailing practices varies around the world. Comparisons of the state of practice worldwide were published by fib
(2003, 2016).
Diaphragm design involves providing load paths through the
floor system [Fig. 9(a)] to carry the in-plane shear, flexure, and
collector forces associated with diaphragm action (Moehle et al.
2010). The joints between the precast units can act as locations
of weakness (for both topped and untopped systems). Thus the
critical regions are often the connections across these joints, which
03118001-8
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
Fig. 8. Precast building applications with walls: (a) Southern Cross Medical Building exterior, Christchurch, New Zealand, postearthquake (image
by Richard S. Henry); (b) building in Chile with unbonded PT walls and hybrid frames (reprinted from Ghosh and Cleland 2012, with permission
from the Earthquake Engineering Research Institute)
Fdia
6.4 102mm
Anchorage
[see Fig. (d)]
10~30m
Collector
Shear
connectors
[see Fig.(b)]
Chord
connectors
[see Fig. (c)]
5~15m
8 216mm
1~2.5m
Shear
wall
102mm
432mm
Beam
or wall
152mm
317mm
#5
U
P
FO
R
1.2~4.6m Slug PL 85x25.4x9.5mm
(a)
8 216mm
(b)
Precast
units
Face plates
216x51x9.6mm
#3
Slug PL 216x38x9.5mm
(c)
(d)
Fig. 9. Precast floor/roof diaphragms (reprinted from Engineering Structures, Vol. 86, Ge Wan, Dichuan Zhang, Robert B. Fleischman, and Clay J.
Naito, “A coupled connector element for nonlinear static pushover analysis of precast concrete diaphragms,“ pp. 58–71, Copyright Elsevier, 2015):
(a) plan; (b) shear connector; (c) chord connector; (d) wall connector
can include mechanical connectors between precast units (standard
industry as well as proprietary connectors), topping slab reinforcement, or a combination of these components [Figs. 9(b–d)].
Precast floor diaphragms have in cases performed poorly in
past earthquakes (EERI 1989, 2010; Iverson and Hawkins 1994;
Saatcioglu et al. 2001; Ghosh and Cleland 2012; Toniolo and
Columbo 2012; Fleischman et al. 2014; Belleri et al. 2014a; Corney
et al. 2014a). These occurrences have limited the use of precast
diaphragms in regions of high seismic hazard in the United States,
whereas more widespread use has been common in other countries
with high seismicity (e.g., New Zealand). In the United States, ACI
318-14 requires a cast-in-place topping over the precast floor units.
UBC-97 (UBC 1997) further limits this usage to a topped noncomposite diaphragm (i.e., the cast-in-place topping slab is designed for
the full diaphragm force transfer, leading to a thicker topping slab
with two-way reinforcement, whereas the precast floor units are
intended to carry gravity loads only), thereby limiting the economic
effectiveness of precast construction.
As such, ongoing research has focused on new designs or details that would permit the use, in high seismic hazard regions, of
© ASCE
untopped diaphragms (Bockemohle 1981; Gates 1981; Menegotto
1994; Zheng 2001; Cleland and Ghosh 2002) or topped composite
diaphragms (i.e., mechanical connectors between precast diaphragm units, acting in conjunction with a thin topping with mesh
or light reinforcing) (Fleischman et al. 2005b; Bull 2004; Fenwick
et al. 2010).
Analytical research on precast diaphragms, which was motivated in large part by past diaphragm failures, has indicated that:
(1) diaphragm design forces can significantly underestimate the inertial forces that develop in the floor system during strong earthquakes (Rodriguez et al. 2007; Fleischman et al. 2002; Priestley
et al. 1999) due to the effect of higher modes during nonlinear
structural response (Rodriguez et al. 2002); (2) nonductile load
paths can develop in diaphragms designed with past design approaches (Wood et al. 2000); (3) diaphragms can possess complicated internal force paths (Clough 1982; Wood et al. 1995; Lee and
Kuchma 2008; Bull 2004), leading to combined tension-shear actions on individual diaphragm connectors (Farrow and Fleischman
2003); (4) nonlinear demands can concentrate at certain key
joints (Fleischman and Wan 2007; Wan et al. 2012); and (5) the
03118001-9
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
diaphragm reinforcement may not possess sufficient nonlinear
deformation capacity (Naito et al. 2009; Cao and Naito 2009)
for these demands. In addition, physical experiments and postearthquake observations have highlighted that axial elongation (from
nonlinear rotation) in beams of moment frames can result in significant nonlinear demands in diaphragms, causing potential
collapse with existing nonductile detailing of support connections
and topping slabs (Fenwick et al. 2005, 2010; Corney et al. 2014a;
Fleischman et al. 2014).
The combination of these factors led to the recognition that current design practice (i.e., diaphragm forces determined from ASCE/
SEI 7-10 Section 12.10.1.1, and possibly the collector forces governed by ASCE/SEI 7-10 12.4.3.2 incorporating the system overstrength factor, Ω0 ) cannot ensure linear-elastic diaphragm action.
Larger diaphragm design forces or improved diaphragm details that
permit reliable nonlinear deformations are needed. Furthermore,
for long-span diaphragms, as often occurs for precast concrete
(e.g., parking structures), the connections can increase diaphragm
flexibility, leading to excessive interstory drifts of the gravity
system columns away from the primary lateral load system
(Fleischman and Farrow 2001). Thus, limits on diaphragm deflections are needed. Nakaki (2000) and Zheng and Oliva (2005)
proposed methods to estimate in-plane diaphragm flexibility,
and Lee et al. (2007) investigated the estimation of interstory drifts
incorporating this flexibility.
The design requirements for diaphragms are being continuously
improved to reflect recent research findings from large-scale subassembly or full structure tests. These tests include simulation-driven
testing of critical precast diaphragm regions [Fig. 10(a)] (Zhang
et al. 2011; Fleischman et al. 2013), cyclic testing of single bay precast frame subassemblies (Dhakal et al. 2014; Fenwick et al. 2010;
Matthews 2004) and a wall system–untopped precast floor subassembly (Liu et al. 2015), pseudodynamic testing of precast buildings
containing diaphragms (Blandón and Rodríguez 2005; Bournas
et al. 2013), and shake table testing [Fig. 10(b)] of diaphragmsensitive precast buildings (Schoettler et al. 2009). In particular,
an industry-supported research program, the diaphragm seismic
design methodology (DSDM) project (Fleischman et al. 2005a,
2013), has resulted in new seismic design recommendations (BSSC
2009; Kelly and Ghosh 2014) and code changes for precast concrete
diaphragms in the United States, as described in more detail
subsequently.
Extensive testing has also been conducted on diaphragm connections (e.g., flange-to-flange connectors, reinforcing across
joints). These include full-scale experiments on isolated precast
connectors (Venuti 1970; Pincheira et al. 1998; Oliva 2000;
Shaikh and Feile 2004; Pang et al. 2012; Watkins et al. 2014), tests
of connected hollow-core panels (Moustafa 1981; Davies et al.
1990) including innovative forms (Menegotto et al. 1998), and tests
of the connection between the floor slab and perimeter framing, or
support connection (Mejia-McMaster and Park 1994; Jensen 2006;
Fenwick et al. 2010; Corney et al. 2014b).
As part of the DSDM project, Naito et al. (2006, 2007) conducted full-scale tests of an assortment of common precast diaphragm connectors under shear, tension, and combined shear and
tension. The purpose of these tests was to prequalify connectors for
use in the new seismic design methodology, emphasizing not only
strength and stiffness, but also reliable nonlinear deformation
capacity. Based on the qualification procedure developed by Naito
and Ren (2013), testing of new and existing connector concepts
have commenced (Naito 2016). Ren and Naito (2013) provide a
comprehensive database of existing tests on precast diaphragm
connectors.
As discussed in more detail subsequently, results from the recent
research on precast diaphragms (Fleischman et al. 2013) are being
adopted into ASCE/SEI 7-16, including an alternative diaphragm
design force calculation for general construction, new diaphragm
design provisions to accompany the new diaphragm design force,
and precast diaphragm connection qualification testing protocols.
The qualification protocols are being developed into an ACI standard for future reference by ACI 318.
The alternative determination of the diaphragm design force
based on rational methods (Rodriguez et al. 2002) can be found
in Section 12.10.3 of ASCE/SEI 7-16. The alternative design force
is calculated using an elastic acceleration coefficient determined by
the square root of sum of squares (SRSS) of the first-mode and
higher-mode responses, in which only the first-mode response is
reduced by the seismic response modification factor, R (and amplified by the overstrength factor, Ωo ). This calculation tends to
produce higher elastic diaphragm forces than previous methods.
However, the alternative design force calculation also includes a
diaphragm force reduction factor, Rs , to account for the inelastic
deformation capacity and overstrength of the diaphragm. In many
cases, the resulting design force may not be significantly different
from that of ASCE/SEI 7-16 Sections 12.10.1 and 12.10.2, particularly if the diaphragm has large inelastic deformation capacity or
overstrength. Detailed explanation of Section 12.10.3 is given in
Part 2 (Commentary) to the 2015 NEHRP Provisions, which is also
Fig. 10. Precast diaphragm tests: (a) panel test (reprinted from Zhang et al. 2011, © ASCE); (b) system test (image courtesy of M. Schoettler,
University of California, San Diego)
© ASCE
03118001-10
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
in the Commentary to ASCE/SEI 7-16. Fleischman et al. (2013)
and DSDM TG (2014) analytically and experimentally verify the
alternative design force level, as described by Ghosh (2016).
ASCE/SEI 7-16 Section 14.2.4 contains specific seismic
design provisions for precast diaphragms and a connector qualification protocol, in addition to applicable seismic design requirements from ASCE/SEI 7-16 and ACI 318-14 Section 18.12. The
precast seismic provisions in Section 14.2.4, based on Zhang and
Fleischman (2016), relate structural parameters and diaphragm
connection classification to the appropriate Rs factor. Diaphragm
connections are classified as high-deformability, moderatedeformability, and low-deformability elements based on testing,
with more deformability permitting a higher Rs factor (and thus
lower design force). The testing qualification protocol follows procedures in Naito et al. (2006, 2007) and Ren and Naito (2013), and
is required to establish the strength, stiffness, and deformation
capacity of any new connector under in-plane shear and tension
(note that certain connections are already prequalified in
Section 14.2.4).
The seismic provisions in Chapter 18 of ACI 318-14 specify
composite as well as noncomposite topped precast diaphragms.
The noncomposite diaphragm is intended for the cast-in-place topping slab to act alone, without the precast floor members contributing to diaphragm force transfer. Reinforcement at the edges of the
cast-in-place topping is considered chord reinforcing to carry the
tension forces due to in-plane flexure. Care must be taken to identify and address configurations where the chord action interacts
with collector or moment frame action in the other orthogonal
direction (DSDM TG 2014). The topping reinforcement over the
floor area provides diaphragm shear transfer. ACI 318-14 Section
18.12.7.4 requires Type 2 mechanical splices when mechanical
splices are used to transfer forces between the diaphragm and the
vertical members of the lateral load-resisting system. Chapter 18
does not include diaphragms with mechanical connections across
the joints between precast units.
Low and moderate seismic risk regions may use untopped double tees relying on mechanical connectors at the joints to transfer
diaphragm in-plane shear (as well as vertical shear), and dry chords
or often cast-in-place pour strips at the ends. The pour strips provide thickened regions in which continuous reinforcement with
larger diameter can be installed to provide the chords. These systems may be designed for high seismic applications on a project-by
project basis (Cleland and Ghosh 2002) until new standards under
development are adopted.
Bridges
This section summarizes the developments associated with using
precast concrete in seismic design of bridges. Generally, research
in this topic was pursued less aggressively until advancements in
accelerated bridge construction took place in nonseismic regions.
Bridge Columns
The preferred location of inelastic action in bridges is at the ends of
columns or piers, thereby minimizing damage to the superstructure
and foundation, and making the inspection and repair of damage
related to inelastic action easier. In emulative construction, the
columns can be connected to precast or cast-in-place bent caps
and foundations through reinforcing bar couplers, grouted ducts,
pockets, and sockets (Fig. 11). Socketed connections can allow
the entire precast column to be embedded in the cap beam or
foundation (Osanai et al. 1996; Haraldsson et al. 2013; Mashal
et al. 2013; Torres Matos and Rodriguez 2014; Belleri and Riva
© ASCE
2012), or with the column longitudinal bars developed in a preformed pocket (Restrepo et al. 2011). Emulative connections with
column bars developed in grouted ducts in the cap beam or foundation have been reported by Pang et al. (2009), Restrepo et al.
(2011), and Tazarv and Saiidi (2015). Bridge column-to-cap-beam
or column-to-foundation connections using grouted splice sleeves
(Ameli et al. 2015) and mechanical splices (Haber et al. 2014) have
also been investigated.
The jointed concept, as demonstrated for building frames and
walls using unbonded or partially debonded PT, has also been investigated for bridge columns. The gap opening and large concentrated
rotations previously described can occur at the column-tofoundation and column-to-cap-beam joints, promoting recentering
response for columns when subjected to seismic loading [Fig. 12(a)].
In most studies discussed in the literature, the column is precast over the entire clear height and connected to the adjoining
member(s) using unbonded PT. As in buildings, special details
are used to avoid or minimize damage at the column-to-foundation
and column-to-cap-beam joints and in the adjacent column critical
regions. To this end, Mander and Cheng (1997) investigated steel
plates at the column-to-foundation joint. Palermo et al. (2007) included steel plates at the top of the foundation and armored the
column toes with steel angles. To ensure satisfactory shear transfer,
they connected a hemispherical steel block to the steel plate.
Tobolski and Restrepo (2008), Cohagen et al. (2008), Restrepo
et al. (2011), Guerrini and Restrepo (2013), and Guerrini et al.
(2014) investigated the types of bedding mortar that can withstand
the impact and transfer shear at the joint between the column and
the cap beam. Tobolski and Restrepo (2008) and Cohagen et al.
(2008) used spirals with a small pitch to confine the column ends.
Billington and Yoon (2004), Trono et al. (2014), and Tazarv and
Saiidi (2015) replaced the column ends with a fiber-reinforced concrete shell, left hollow or filled with self-consolidating concrete.
Tobolski and Restrepo (2008), Restrepo et al. (2011), Guerrini
and Restrepo (2013), and Guerrini et al. (2014) investigated columns built with a dual steel shell (with concrete cast in between
the shells) designed for composite action. Guerrini and Restrepo
(2013), Guerrini et al. (2014), and Trono et al. (2014) used headed
reinforcing bars at the column end with matching headed bars embedded in the foundation to transfer compression. ElGawady and
Sha’lan (2010) used thin neoprene pads at the column ends and
found that the sheets significantly reduced the lateral stiffness of
the column. Motaref et al. (2014) used a laminated elastomeric
bearing at the column ends, where typically a plastic hinge would
develop in a conventional column. The use of these materials will
alter the dynamic characteristics of the column, which should be
given attention in determining the expected response. Furthermore,
because these materials are Voigt-type, the strain rate effect may be
significant and should be included when determining the dynamic
response of the column.
Large columns or columns where the lifting capacity is limited
can be segmented and then posttensioned using unbonded tendons.
Hewes and Priestley (2002), Billington and Yoon (2004), Chou and
Chen (2006), Shim et al. (2008), Wang et al. (2008), Ou et al. (2007,
2009, 2010), Taira et al. (2009), Yamashita and Sanders (2009),
Sakai et al. (2009), Kim et al. (2010a, b, 2012), Motaref et al.
(2014), and Dawood et al. (2012) investigated the response of precast segmental columns where all joints could undergo gap opening,
allowing the column to rock, while preventing (or minimizing) shear
sliding. Sideris et al. (2014b, 2015) investigated a precast segmental
method where nonlinear deformations could occur either due to
rocking or through sliding.
Mashal and Palermo (2015) and White and Palermo (2016) investigated unbonded PT bridge columns embedded in a socket in
03118001-11
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
Fig. 11. Different types of precast bridge column connections (reprinted from Marsh et al. 2011, with permission from the Transportation Research
Board): (a) rebar coupler connection; (b) grouted connection; (c) pocket connection; (d) socket connection
(a)
(b)
Fig. 12. Precast bridges: (a) kinematics of a bridge bent using unbonded PT columns (reprinted from Guerrini et al. 2014, © ASCE); (b) segmental
bridge construction—Otay Mesa Bridge in California (image courtesy of International Bridge Technologies, Inc.)
the foundation and armored with a steel jacket. Davis et al. (2012)
and Eberhard et al. (2014) demonstrated that precast pretensioned
bridge columns incorporating partially debonded strands can also
be used to ensure recentering. In these alternatives, the critical region at the column ends was confined with either spirals (Davis
et al. 2012) or a hybrid fiber-reinforced concrete (HyFRC) shell
(Eberhard et al. 2014).
© ASCE
Supplementary energy dissipation has been provided to jointed
bridge columns by partially debonded mild or stainless-steel
reinforcing bars (Cohagen et al. 2008; Tobolski and Restrepo
2008; Ou et al. 2010; Restrepo et al. 2011; Guerrini and
Restrepo 2013; Guerrini et al. 2014; Eberhard et al. 2014;
Thonstad et al. 2016) or replaceable bars (Tazarv and Saiidi
2015; White and Palermo 2016). Marriot et al. (2009, 2011) and
03118001-12
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Guerrini et al. (2014) investigated replaceable external hysteretic
dampers.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
Bridge Superstructure
Precast bridge superstructures include bent cap beams, girders, and
decks. The superstructure, including the cap beams, is typically
capacity-designed to remain essentially linear-elastic under the
column overstrength moment resistance. This can be successfully
accomplished using emulative and/or jointed connections between
the bridge superstructure members. The effect of vertical acceleration should be given consideration when establishing the connection demands.
Only a few studies have focused on the seismic response of precast bridge superstructures and their connections. Sritharan et al.
(2001) investigated the use of precast cap beams in multicolumn
bents, and Holombo et al. (2000) investigated the seismic response
of spliced precast I-girders and bathtub girders. Restrepo et al.
(2011) and Vander Werff et al. (2015) reported the response of
two different precast I-girder-to-cap-beam connections. In box
girder bridges, large precast segments connected with bonded
and/or external (unbonded) PT can be used along the bridge length
[Fig. 12(b)]. Megally et al. (2002, 2003a, b) and Veletzos and
Restrepo (2009, 2010, 2011) examined the seismic behavior of precast segmental bridge superstructures with emphasis on gap opening at the joints. The joints between the superstructure segments, as
well as those between the superstructure and the column, were also
investigated under quasi-static testing by Burnell et al. (2004) and
by multiaxial shake table excitation by Sideris et al. (2015), who
performed a system test and included precast segmental columns.
Summary, Concluding Remarks, and Future
Directions
A significant amount of research related to precast seismic
structural systems has been conducted in the United States and elsewhere since the 1980s. The early part of this research focused on
the development of emulative connections (between precast structural members and with the foundation), which were aimed at ensuring that the seismic behavior of precast structures would be
comparable to that of equivalent monolithic cast-in-place reinforced concrete structures. Since the introduction of jointed connections for critical regions, the focus has shifted toward this
connection concept because it takes advantage of unique properties
of precast concrete and offers better-performing structures than do
comparable alternatives. This paper serves as a resource document
for researchers and practitioners by providing an overview of related research and summarizing the important advances on the
use of precast concrete in moment frames, structural walls, floor
diaphragms, and bridges. It also discusses accompanying changes
to governing design codes and standards, and selected practical implementations of seismic precast structures.
This review of the accomplishments in research and practice
makes clear that the use of precast concrete has been significantly
advanced and that code-compliant precast structures can be designed today in regions of high seismic hazard. These structures
can not only be cost effective and of high quality due to the use
of precast concrete, but they will also produce better seismic response and be functional following a major seismic event when
appropriate connection details are used in the critical regions. These
benefits are significant, and are likely to increase the use of precast
concrete construction in the years to come.
Although tremendous research progress has been made toward
the use of precast concrete in high seismic regions, the transfer of
© ASCE
this progress for application in practice has been relatively slow.
Among the jointed systems, the hybrid moment frame has been
used more extensively than any other precast concept in seismic
regions. The relatively slow practical adoption of seismic precast
systems may be attributed to concerns associated with the anchorages of unbonded PT steel and concerns associated with the interaction of the lateral load resisting system (e.g., walls) with other
parts of the structure (e.g., gravity columns and floor/roof diaphragms). Further research to overcome these concerns and further
development/improvement of design guidelines and standards will
take place in coming years. These efforts will contribute to the development of confidence for the widespread use of precast systems
in high seismic regions in the future.
References
Aaleti, S., and Sritharan, S. (2009). “A simplified analysis method for characterizing unbonded post-tensioned precast wall systems.” Eng. Struct.,
31(12), 2966–2975.
ACI (American Concrete Institute). (2001). “Design recommendations
for precast concrete structures (reapproved 2001).” ACI 550R-96,
Farmington Hills, MI.
ACI (American Concrete Institute). (2003). “Special hybrid moment frames
composed of discretely jointed precast and post-tensioned conc.
members.” ACI ITG T1.2-03/T1.2R-03, Farmington Hills, MI, 15.
ACI (American Concrete Institute). (2005). “Acceptance criteria for moment frames based on structural testing and commentary (reapproved
2014).” ACI 374.1-05, Farmington Hills, MI.
ACI (American Concrete Institute). (2007). “Acceptance criteria for special
unbonded post-tensioned precast structural walls based on validation
testing and commentary.” ACI ITG 5.1-07, Farmington Hills, MI.
ACI (American Concrete Institute). (2008). “Building code requirements
for structural concrete and commentary.” ACI Committee 318-08,
Farmington Hills, MI.
ACI (American Concrete Institute). (2009). “Requirements for design of
a special unbonded post-tensioned precast shear wall satisfying ACI
ITG-5.1 and commentary.” ACI ITG 5.2-09, Farmington Hills, MI.
ACI (American Concrete Institute). (2014). “Building code requirements
for structural concrete and commentary.” ACI Committee 318-14,
Farmington Hills, MI.
AIJ (Architectural Institute of Japan). (2000). Draft of Japanese design
guidelines for precast construction of equivalent monolithic reinforced
concrete buildings, Tokyo.
Ajrab, J., Pekcan, G., and Mander, J. (2004). “Rocking wall-frame structures with supplemental tendon systems.” J. Struct. Eng., 10.1061
/(ASCE)0733-9445(2004)130:6(895), 895–903.
Allen, M., and Kurama, Y. (2002a). “Design of rectangular openings
in precast walls under combined vertical and lateral loads.” PCI J.,
47(2), 58–83.
Allen, M., and Kurama, Y. (2002b). “Design of rectangular openings in
precast walls under vertical loads.” PCI J., 47(1), 50–67.
Ameli, M. J., Parks, J. E., Brown, D. N., and Pantelides, C. P. (2015).
“Seismic evaluation of grouted splice sleeve connections for reinforced
precast concrete column-to-cap beam joints in accelerated bridge
construction.” PCI J., 60(2), 80–103.
Aragon, T., Kurama, Y., and Meinheit, D. (2017). “A Type III grouted
seismic connector for precast concrete structures.” PCI J., 62(5),
75–88.
Armouti, N. (1993). “Seismic performance of precast concrete structural
walls.” Ph.D. dissertation, Lehigh Univ., Bethlehem, PA.
ASCE. (2013). “Minimum design loads for buildings and other structures.”
ASCE/SEI 7-10, Reston, VA.
ASCE. (2016). “Minimum design loads for buildings and other structures.”
ASCE/SEI 7-16, Reston, VA.
Belleri, A., Brunesi, E., Nascimbene, R., Pagani, M., and Riva, P.
(2014a). “Seismic performance of precast industrial facilities following
major earthquakes in the Italian territory.” J. Perform. Constr. Facil.,
10.1061/(ASCE)CF.1943-5509.0000617, 04014135.
03118001-13
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
Belleri, A., and Riva, P. (2012). “Seismic performance and retrofit of precast concrete grouted sleeve connections.” PCI J., 57(1), 97–109.
Belleri, A., Schoettler, M. J., Restrepo, J. I., and Fleischman, R. B. (2014b).
“Dynamic behavior of rocking and hybrid cantilever walls in a precast
concrete building.” ACI Struct. J., 111(3), 661.
Billington, S., and Yoon, J. (2004). “Cyclic response of unbonded posttensioned precast columns with ductile fiber-reinforced concrete.”
J. Bridge Eng., 10.1061/(ASCE)1084-0702(2004)9:4(353), 353–363.
Blandón, J. J., and Rodríguez, M. E. (2005). “Behavior of connections
and floor diaphragms in seismic-resisting precast concrete buildings.”
PCI J., 50(2), 56–75.
Bockemohle, L. W. (1981). “A practical paper on design of topped concrete
diaphragms and precast concrete structures.” Workshop on Design of
Prefabricated Concrete Buildings for Earthquake Loads, Applied
Technology Council, Redwood City, CA.
Bournas, D. A., Negro, P., and Molina, F. J. (2013). “Pseudodynamic tests
on a full-scale 3-storey precast concrete building: Behavior of the
mechanical connections and floor diaphragms.” Eng. Struct., 57,
609–627.
BSSC (Building Seismic Safety Council). (2001). NEHRP recommended
provisions for seismic regulations for new buildings and other structures, Washington, DC.
BSSC (Building Seismic Safety Council). (2009). Seismic design methodology for precast concrete floor diaphragms, Washington, DC.
BSSC (Building Seismic Safety Council). (2015). NEHRP recommended
seismic provisions for new buildings and other structures, Washington,
DC.
Bull, D. K. (2004). “Understanding the complexities of designing diaphragms in buildings for earthquakes.” Bull. N. Z. Soc. Earthquake
Eng., 37(2), 70–88.
Burnell, K., Restrepo, J. I., and Megally, S. (2004). “Longitudinal testing
of a precast post-tensioned bridge system.” 20th US-Japan Workshop
on Bridge Engineering, Public Works Research Institute of Japan,
Washington, DC.
CAE (Centre for Advanced Engineering). (1999). Guidelines for the use of
structural precast concrete in buildings, Christchurch, New Zealand.
Caltrans. (2013). Seismic design criteria, version 1.7, California Dept. of
Transportation, Sacramento, CA.
Cao, L., and Naito, C. (2009). “Precast double-tee floor connectors. Part 2:
Shear performance.” PCI J., 54(2), 97–115.
Cheok, G., and Lew, H. (1993). “Model precast concrete beam-to-column
connections subject to cyclic loading.” PCI J., 38(4), 80–92.
Chou, C. C., and Chen, Y. C. (2006). “Cyclic tests of post-tensioned precast
CFT segmental bridge columns with unbonded strands.” Earthquake
Eng. Struct. Dyn., 35(2), 159–175.
Cleland, N. M., and Ghosh, S. K. (2002). “Untopped precast concrete diaphragms in high-seismic applications.” PCI J., 47(6), 94–99.
Cleland, N. M., and Ghosh, S. K. (2012). Seismic design of precast/
prestressed concrete structures, Precast/Prestressed Concrete Institute,
Chicago.
Clough, D. P. (1982). “Considerations in the design and construction of
precast concrete diaphragms for earthquake loads.” PCI J., 27(2),
79–93.
Cohagen, L. S., Pang, J. B. K., Eberhard, M. O., and Stanton, J. F. (2008).
“A precast concrete bridge bent designed to recenter after an earthquake.” Rep. No. WA-RD 684.3, Washington State Transportation
Center, Seattle.
Corney, S. R., Henry, R. S., and Ingham, J. M. (2014a). “Performance of
precast concrete floor systems during the 2010/2011 Canterbury earthquake series.” Mag. Concr. Res., 66(11), 563–575.
Corney, S. R., Henry, R. S., and Ingham, J. M. (2014b). “Testing support
connections of 400 hollow-core precast concrete floors.” 10th U.S.
National Conf. on Earthquake Engineering, Earthquake Engineering
Research Institute, Oakland, CA.
Crisafulli, F. J., Restrepo, J. I., and Park, R. (2002). “Seismic design of
lightly reinforced precast concrete rectangular wall panels.” PCI J.,
47(4), 104–121.
D’Arcy, T. J., Nasser, G. D., and Ghosh, S. K. (2003). “Building code provisions for precast/prestressed concrete: A brief history.” PCI J., 48(6),
116–124.
© ASCE
Davies, G., Elliott, K. S., and Omar, W. (1990). “Horizontal diaphragm
action in precast concrete floors.” Struct. Eng., 68(2), 25–33.
Davis, P. M., Janes, T. M., Eberhard, M. O., and Stanton, J. F. (2012).
“Unbonded pre-tensioned columns for bridges in seismic regions.”
PEER Rep. 2012/04, Pacific Earthquake Engineering Research Center,
Berkeley, CA.
Dawood, H., ElGawady, M., and Hewes, J. (2012). “Behavior of segmental
precast posttensioned bridge piers under lateral loads.” J. Bridge Eng.,
10.1061/(ASCE)BE.1943-5592.0000252, 735–746.
Dhakal, R. P., Peng, B. H. H., Fenwick, R. C., Carr, A. J., and Bull, D. K.
(2014). “Cyclic loading test of reinforced concrete frame with precastprestressed flooring system.” ACI Struct. J., 111(4), 777–788.
DSDM TG (Diaphragm Seismic Design Methodology Task Group). (2014).
“Development and design of untopped precast concrete diaphragm
system for high seismic zones.” ⟨http://www.pankowfoundation.org
/grants.cfm?grantid=1070⟩ (Feb. 16, 2014).
Eberhard, M. O., Stanton, J. F., Haraldsson, O. S., Finnsson, G., Davis,
P. M., and Schoettler, M. J. (2014). “Development of a bridge bent
system for rapid construction and enhanced seismic performance.”
10th U.S. National Conf. on Earthquake Engineering, Earthquake
Engineering Research Institute, Oakland, CA.
EERI (Earthquake Engineering Research Institute). (1989). Armenia earthquake reconnaissance report, Oakland, CA.
EERI (Earthquake Engineering Research Institute). (2010). “The El Mayor
Cucapah, Baja California earthquake.” An EERI learning from earthquakes reconnaissance report, J. Meneses, ed., Oakland, CA.
ElGawady, M. A., and Sha’lan, A. (2010). “Seismic behavior of selfcentering precast segmental bridge bents.” J. Bridge Eng., 10.1061
/(ASCE)BE.1943-5592.0000174, 328–339.
El-Sheikh, M. T., Sause, R., Pessiki, S., and Lu, L.-W. (1999). “Seismic
behavior and design of unbonded post-tensioned precast concrete
frames.” PCI J., 44(3), 54–71.
Englekirk, R. (1986). “Overview of PCI workshop on effective use of
precast concrete for seismic resistance.” PCI J., 31(6), 48–58.
Englekirk, R. (2002). “Design-construction of the paramount: A 39-story
precast prestressed concrete apartment building.” PCI J., 47(4), 56–71.
Ericson, A. C., and Warnes, C. E. (1990). Seismic technology for precast
concrete systems, Concrete Industry Board, Inc., Kew Gardens, NY.
Erkmen, B., and Schultz, A. E. (2009). “Self-centering behavior of unbonded, post-tensioned precast concrete shear walls.” J. Earthquake
Eng., 13(7), 1047–1064.
Farrow, K. T., and Fleischman, R. B. (2003). “Effect of dimension and detail on the capacity of precast concrete parking structure diaphragms.”
PCI J., 48(5), 46–61.
Farrow, K. T., and Kurama, Y. C. (2003). “SDOF demand index relationships for performance-based seismic design.” Earthquake Spectra,
19(4), 799–838.
Fenwick, R., Bull, D., and Gardiner, D. (2010). “Assessment of hollowcore floors for seismic performance.” Rep. No. 2010-02, Univ. of
Canterbury, Christchurch, New Zealand.
Fenwick, R. C., Davidson, B. J., and Lau, D. B. N. (2005). “Interaction
between ductile RC perimeter frames and floor slabs containing precast
units.” New Zealand Society for Earthquake Engineering Conf., New
Zealand Society for Earthquake Engineering, Wellington, New Zealand.
fib (Fédération internationale du béton). (2003). Seismic design of precast
concrete buildings, Lausanne, Switzerland, 262.
fib (Fédération internationale du béton). (2016). Pre-cast concrete buildings in seismic areas, Lausanne, Switzerland, 273.
Fleischman, R. B., et al. (2005a). “Seismic design methodology for precast concrete diaphragms. Part 2: Research program.” PCI J., 51(6),
2–19.
Fleischman, R. B., Farrow, K., and Eastman, K. (2002). “Seismic response
of perimeter lateral-system structures with highly flexible diaphragms.”
Earthquake Spectra, 18(2), 251–286.
Fleischman, R. B., and Farrow, K. T. (2001). “Dynamic response of perimeter lateral-system structures with flexible diaphragms.” Earthquake
Eng. Struct. Dyn., 30(5), 745–763.
Fleischman, R. B., Naito, C. J., Restrepo, J., Sause, R., and Ghosh, S. K.
(2005b). “Seismic design methodology for precast concrete diaphragms
Part 1: Design framework.” PCI J., 50(5), 68–83.
03118001-14
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
Fleischman, R. B., Restrepo, J. I., Naito, C., Sause, R., Zhang, D., and
Schoettler, M. (2013). “Integrated analytical and experimental research
to develop a new seismic design methodology for precast concrete diaphragms.” J. Struct. Eng., 10.1061/(ASCE)ST.1943-541X.0000734,
1192–1204.
Fleischman, R. B., Restrepo, J. I., Pampanin, S., Maffei, J. R., Seeber, K.,
and Zahn, F. A. (2014). “Damage evaluations of precast concrete structures in the 2010–2011 Canterbury earthquake sequence.” Earthquake
Spectra, 30(1), 277–306.
Fleischman, R. B., and Wan, G. (2007). “Appropriate overstrength of shear
reinforcement in precast concrete diaphragms.” J. Struct. Eng., 10.1061
/(ASCE)0733-9445(2007)133:11(1616), 1616–1626.
Galusha, J. G. (1999). “Precast, post-tensioned concrete walls designed to
rock.” M.S.C.E. thesis, Univ. of Washington, Seattle.
Gates, W. E. (1981). “Seismic design considering for untopped precast
concrete floor and roof diaphragm.” Workshop on Design of Prefabricated Concrete Buildings for Earthquake Loads, Applied Technology
Council, Redwood City, CA.
Ghosh, S. K. (2004). “Update on the NEHRP provisions: The resource
document for seismic design.” PCI J., 49(3), 96–102.
Ghosh, S. K. (2016). “Alternative diaphragm seismic design force level of
ASCE 7-16.” Structure Magazine, Mar., 18–23.
Ghosh, S. K., and Cleland, N. M. (2012). “Performance of precast concrete
building structures.” Earthquake Spectra, 28(S1), S349–S384.
Guerrini, G., and Restrepo, J. (2013). “Seismic response of composite
concrete-dual steel shell columns for accelerated bridge construction.”
7th National Seismic Conf. on Bridges and Highways, State Univ. of
New York, Buffalo, NY.
Guerrini, G., Restrepo, J. I., Massari, M., and Vervelidis, A. (2014).
“Seismic behavior of posttensioned self-centering precast concrete
dual-shell steel columns.” J. Struct. Eng., 10.1061/(ASCE)ST.1943541X.0001054, 04014115.
Haber, Z. B., Saiidi, M. S., and Sanders, D. H. (2014). “Seismic performance of precast columns with mechanically spliced column-footing
connections.” ACI Struct. J., 111(3), 639.
Haraldsson, O. S., Janes, T. M., Eberhard, M. O., and Stanton, J. F. (2013).
“Seismic resistance of socket connection between footing and precast
column.” J. Bridge Eng., 10.1061/(ASCE)BE.1943-5592.0000413,
910–919.
Hawkins, N. M., and Englekirk, R. (1987). “U.S.-Japan seminar on precast
concrete construction in seismic zones.” PCI J., 32(2), 75–85.
Hawkins, N. M., and Ghosh, S. K. (2004). “Acceptance criteria for special
precast concrete structural walls based on validation testing.” PCI J.,
49(5), 78–92.
Hawkins, N. M., and Ghosh, S. K. (2006). “Codification of precast seismic
structural systems: An update.” PCI J., 51(3), 46–49.
Henry, R. S., Brooke, N. J., Sritharan, S., and Ingham, J. M. (2012).
“Defining concrete compressive strain in unbonded post-tensioned
walls.” ACI Struct. J., 109(1), 101–112.
Hewes, J. T., and Priestley, M. N. (2002). “Seismic design and performance
of precast concrete segmental bridge columns.” Rep. No. SSRP-2001/25,
Univ. of California, San Diego.
Holden, T., Restrepo, J., and Mander, J. (2001). “A comparison of the seismic performance of precast wall construction: Emulation and hybrid
approaches.” Rep. No. 2001-4, Univ. of Canterbury, Christchurch,
New Zealand.
Holden, T., Restrepo, J., and Mander, J. B. (2003). “Seismic performance
of precast reinforced and prestressed concrete walls.” J. Struct. Eng.,
10.1061/(ASCE)0733-9445(2003)129:3(286), 286–296.
Holombo, J. M., Priestley, J. N., and Seible, F. (2000). “Continuity of precast prestressed spliced-girder bridges under seismic loads.” PCI J.,
45(2), 40–63.
Housner, G. and Lili, X. (2002). The great Tangshan earthquake of 1976:
Overview volume, G. W. Housner and H. Duxin, eds., Earthquake
Engineering Laboratory, California Institute of Technology, Pasadena,
CA.
IBC (International Building Code). (2014). “2015 international building
code.” IBC-15, International Code Council, Country Club Hills, IL.
© ASCE
Iverson, J., and Hawkins, N. (1994). “Performance of precast/prestressed
concrete building structures during Northridge earthquake.” PCI J.,
39(2), 38–55.
Jensen, J. (2006). “Experimental investigation of existing hollow-core seating connections pre and post retrofit.” M.E. thesis, Univ. of Canterbury,
Christchurch, New Zealand.
Kelly, D. J., and Ghosh, S. K. (2014). “Diaphragm response to and design
for earthquake ground motions.” 10th U.S. National Conf. on
Earthquake Engineering, Earthquake Engineering Research Institute,
Oakland, CA.
Kim, T.-H., Lee, H.-M., Kim, Y.-J., and Shin, H. M. (2010a). “Performance
assessment of precast concrete segmental bridge columns with a shear
resistant connecting structure.” Eng. Struct., 32(5), 1292–1303.
Kim, T.-H., Park, S.-J., Kim, Y-J., and Shin, H. M. (2010b). “Performance
assessment of precast segmental PSC bridge columns with precast concrete footings.” Mag. Concr. Res., 62(11), 773–787.
Kim, T.-H., Seong, D.-J., and Shin, H. M. (2012). “Seismic performance
assessment of hollow reinforced concrete and prestressed concrete
bridge columns.” Int. J. Concrete Struct. Mater., 6(3) 165–176.
Kurama, Y. (2000). “Seismic design of unbonded post-tensioned precast
walls with supplemental viscous damping.” ACI Struct. J., 97(4),
648–658.
Kurama, Y. (2001). “Simplified seismic design approach for frictiondamped unbonded post-tensioned precast walls.” ACI Struct. J.,
98(5), 705–716.
Kurama, Y. (2002). “Hybrid post-tensioned precast concrete walls for use in
seismic regions.” PCI J., 47(5), 36–59.
Kurama, Y. (2005). “Seismic design of partially post-tensioned precast
concrete walls.” PCI J., 50(4), 100–125.
Kurama, Y., Pessiki, S., Sause, R., and Lu, L. W. (1999a). “Seismic behavior and design of unbonded post-tensioned precast concrete walls.”
PCI J., 44(3), 72–89.
Kurama, Y., Sause, R., Pessiki, S., and Lu, L. W. (1999b). “Lateral load
behavior and seismic design of unbonded post-tensioned precast concrete walls.” ACI Struct. J., 96(4), 622–632.
Kurama, Y., Sause, R., Pessiki, S., and Lu, L. W. (2002). “Seismic response
evaluation of unbonded post-tensioned precast walls.” ACI Struct. J.,
99(5), 641–651.
Lee, H. J., Aschheim, M. A., and Kuchma, D. (2007). “Interstory drift estimates for low-rise flexible diaphragm structures.” Eng. Struct., 29(7),
1375–1397.
Lee, H. J., and Kuchma, D. A. (2008). “Seismic response of parking structures with precast concrete diaphragms.” PCI J., 53(2), 71–94.
Liu, Q., Watkins, J., French, C., Sritharan, S., and Nakaki, S. (2015).
Rocking wall-floor-column subassemblage pseudo-static cyclic test at
NEES@UMN (MAST) specimen 2 [Data set], Network for Earthquake
Engineering Simulation, Austin, TX.
Mander, J. B., and Cheng, C-T. (1997). “Seismic resistance of bridge piers
based on damage avoidance design.” Technical Rep. NCEER-97-0014,
State Univ. of New York, Buffalo, NY, 109.
Marriott, D., Pampanin, S., Bull, D., and Palermo, A. (2008). “Dynamic
testing of precast, post-tensioned rocking wall systems with alternative dissipating solutions.” Bull. N. Z. Soc. Earthquake Eng., 41(2),
90–103.
Marriott, D., Pampanin, S., and Palermo, A. (2009). “Quasi-static and
pseudo-dynamic testing of unbonded post-tensioned rocking bridge
piers with external replaceable dissipaters.” Earthquake Eng. Struct.
Dyn., 38(3), 331–354.
Marriott, D., Pampanin, S., and Palermo, A. (2011). “Biaxial testing of unbonded post-tensioned rocking bridge piers with external replaceable
dissipaters.” Earthquake Eng. Struct. Dyn., 40(15), 1723–1741.
Marsh, M. L., Wernli, M., Garrett, B. E., Stanton, J. F., Eberhard, M. O.,
and Weinert, M. D. (2011). “Application of accelerated bridge construction connections in moderate-to-high seismic regions.” NCHRP Rep.
698, Transportation Research Board of the National Academies,
Washington, DC, 54.
Martin, L., and Korkosz, W. (1982). “Connections of precast prestressed
concrete buildings, including earthquake resistance.” Technical Rep.
No. 2, Prestressed Concrete Institute, Chicago.
03118001-15
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
Mashal, M., and Palermo, A. (2015). “High-damage and low-damage
seismic design technologies for accelerated bridge construction.” ASCE
Structures Congress, ASCE, Portland, OR, 549–560.
Mashal, M., White, S., and Palermo, A. (2013). “Quasi-static cyclic tests of
emulative precast segmental bridge piers (E-PSBP).” New Zealand
Society for Earthquake Engineering Conf., New Zealand Society for
Earthquake Engineering, Wellington, New Zealand.
Matthews, J. (2004). “Hollow-core floor slab performance following a
severe earthquake.” Ph.D. thesis, Univ. of Canterbury, Christchurch,
New Zealand.
Megally, S., Seible, F., and Dowell, R. K. (2003a). “Seismic performance of
precast segmental bridges: Segment-to-segment joints subjected to high
flexural moments and high shears.” PCI J., 48(3), 72–90.
Megally, S., Seible, F., and Dowell, R. K. (2003b). “Seismic performance
of precast segmental bridges: Segment-to-segment joints subjected to
high flexural moments and low shears.” PCI J., 48(2), 80–96.
Megally, S., Seible, F., Garg, M., and Dowell, R. K. (2002). “Seismic performance of precast segmental bridge superstructures with internally
bonded prestressing tendons.” PCI J., 47(2), 40–56.
Mejia-McMaster, J. C., and Park, R. (1994). “Tests on special reinforcement for end support of hollow-core slabs.” PCI J., 39(5), 90–105.
Menegotto, M. (1994). “Seismic diaphragm behavior of untopped hollowcore floors.” 12th FIP Congress, Int. Federation for Prestressing,
Federation Internationale de la Precontrainte, Lausanne, Switzerland,
E3–E9.
Menegotto, M., Monti, G., and Auriccio, F. (1998). “Diaphragm modeling
of precast floors with frictional joints.” 11th European Conf. on Earthquake Engineering, Bogaziçi Univ., Istanbul, Turkey.
Moehle, J. P., Hooper, J., Kelly, D., and Meyer, T. (2010). “Seismic design
of cast-in-place concrete diaphragms, chords, and collections.” NIST
GCR 10-917-4, National Institute of Standards and Technology,
Gaithersburg, MD.
Morgen, B., and Kurama, Y. (2004). “A friction damper for post-tensioned
precast concrete moment frames.” PCI J., 49(4), 112–133.
Morgen, B., and Kurama, Y. (2007). “Seismic design of friction-damped
precast concrete frame structures.” J. Struct. Eng., 10.1061/(ASCE)
0733-9445(2007)133:11(1501), 1501–1511.
Morgen, B., and Kurama, Y. (2008). “Seismic response evaluation of posttensioned precast concrete frames with friction dampers.” J. Struct.
Eng., 10.1061/(ASCE)0733-9445(2008)134:1(132), 132–145.
Motaref, S., Saiidi, M. S., and Sanders, D. (2014). “Shake table studies of
energy-dissipating segmental bridge columns.” J. Bridge Eng., 10.1061
/(ASCE)BE.1943-5592.0000518, 186–199.
Moustafa, S. E. (1981). “Effectiveness of shear-friction reinforcement
in shear diaphragm capacity of hollow-core slabs.” PCI J., 26(1),
118–133.
Muguruma, H., Nishiyama, M., and Watanabe, F. (1995). “Lessons from
the Kobe earthquake: A Japanese perspective.” PCI J., 40(4), 28–42.
Muir, C. A., Bull, D. K., and Pampanin, S. (2012). “Preliminary observations from biaxial testing of a two-storey, two-by-one bay, reinforced
concrete slotted beam superassembly.” Bull. N. Z. Soc. Earthquake
Eng., 45(3), 97–104.
Musselman, E., Fournier, M., McAlpine, P., and Sritharan, S. (2015).
“Behavior of unbonded post-tensioning monostrand anchorage systems
under short duration, high amplitude cyclical loading.” Eng. Struct.,
104, 116–125.
Nagae, T., et al. (2014). “The 2010 E-defense shaking table test on fourstory reinforced concrete and post-tensioned concrete buildings.”
10th U.S. National Conf. on Earthquake Engineering: Frontiers of
Earthquake Engineering, Earthquake Engineering Research Institute,
Oakland, CA.
Naito, C. (2016). “In-plane qualification of the meadow burke seismic
chord connector.” ATLSS Rep. No. 16-02, Lehigh Univ., Bethlehem,
PA, 22.
Naito, C., Cao, L., and Peter, W. (2009). “Precast double-tee floor connections. Part 1: Tension performance.” PCI J., 54(1), 49–66.
Naito, C., Jones, C., Cullen, T., and Ren, R. (2007). “Development of a
seismic design methodology for precast diaphragms—Phase 1b summary report.” ATLSS Rep., Lehigh Univ., Bethlehem, PA.
© ASCE
Naito, C., Peter, W., and Cao, L. (2006). “Development of a seismic design
methodology for precast diaphragms—Phase 1 summary report.”
ATLSS Rep. No. 06-01, Lehigh Univ., Bethlehem, PA.
Naito, C., and Ren, R. (2013). “An evaluation method for precast concrete
diaphragm connectors based on structural testing.” PCI J., 58(2),
106–118.
Nakaki, S. D. (2000). “Design guidelines for precast and cast-in-place concrete diaphragms.” Technical Rep., Earthquake Engineering Research
Institute, Oakland, CA.
Nakaki, S. D., Stanton, J., and Sritharan, S. (1999). “An overview of the
PRESSS five-story precast test building.” PCI J., 44(2), 26–39.
Nazari, M., Sritharan, S., and Aaleti, A. (2017). “Single precast concrete
rocking walls as earthquake force-resisting elements.” Earthquake Eng.
Struct. Dyn., 46(5), 753–769.
Negro, P., Bournas, D. A., and Molina, F. J. (2013). “Pseudodynamic tests
on a full-scale 3-storey precast concrete building: Global response.”
Eng. Struct., 57, 594–608.
Nishiyama, M. (1990). “Seismic design of prestressed concrete buildings.”
Bull. N. Z. Natl. Soc. Earthquake Eng., 23(4), 288–304.
Oliva, M. (2000). Testing of the JVI flange connector for precast concrete
double-tee system, Univ. of Wisconsin, Madison, WI.
Oliva, M., Clough, R., and Malhas, F. (1989). “Seismic behavior of large
panel precast concrete walls: Analysis and experiment.” PCI J., 34(5),
42–66.
Oliva, M., Gavrilovic, P., and Clough, R. (1990). “Seismic testing of large
panel precast walls: Comparison of pseudostatic and shaking table
tests.” Earthquake Eng. Struct. Dyn., 19(6), 859–875.
Osanai, Y., Watanabe, F., and Okamoto, S. (1996). “Stress transfer mechanism of socket base connections with precast concrete columns.”
ACI Struct. J., 93(3), 266–276.
Ou, Y. C., Chiewanichakorn, M., Aref, A. J., and Lee, G. C. (2007).
“Seismic performance of segmental precast unbonded posttensioned
concrete bridge columns.” J. Struct. Eng., 10.1061/(ASCE)0733-9445
(2007)133:11(1636), 1636–1647.
Ou, Y. C., Tsai, M. S., Chang, K. C., and Lee, G. C. (2010). “Cyclic behavior of precast segmental concrete bridge columns with high performance or conventional steel reinforcing bars as energy dissipation
bars.” Earthquake Eng. Struct. Dyn., 39(11), 1181–1198.
Ou, Y. C., Wang, P. H., Tsai, M. S., Chang, K. C., and Lee, G. C. (2009).
“Large-scale experimental study of precast segmental unbonded posttensioned concrete bridge columns for seismic regions.” J. Struct. Eng.,
10.1061/(ASCE)ST.1943-541X.0000110, 255–264.
Palermo, A., Pampanin, S., and Marriott, D. (2007). “Design, modeling,
and experimental response of seismic resistant bridge piers with posttensioned dissipating connections.” J. Struct. Eng., 10.1061/(ASCE)
0733-9445(2007)133:11(1648), 1648–1661.
Pampanin, S., Kam, W. Y., Haverland, G., and Gardiner, S. (2011).
“Expectation meets reality: Seismic performance of a post-tensioned
precast concrete building (PRESSS technology) during the 22nd Feb
2011 Christchurch Earthquake.” New Zealand Concrete Industry Conf.,
New Zealand Concrete Society, Auckland, New Zealand.
Pampanin, S., Marriott, D., and Palermo, A. (2010). PRESSS design handbook, New Zealand Concrete Society, Auckland, New Zealand.
Pang, J. B., Eberhard, M. O., and Stanton, J. F. (2009). “Large-bar connection for precast bridge bents in seismic regions.” J. Bridge Eng.,
10.1061/(ASCE)BE.1943-5592.0000081, 231–239.
Pang, R., Liang, S., and Zhu, X. (2012). “Experimental study on seismic
behavior of slab connections in untopped precast RC floor diaphragms.”
J. Build. Struct., 33(10), 59–66.
Park, R. (1990). “Precast concrete in seismic resisting buildings in
New Zealand.” Concr. Int., 12(11), 43–51.
Park, R. (1995). “A perspective on the seismic design of precast concrete
structures in New Zealand.” PCI J., 40(3), 40–60.
Park, R., and Bull, D. K. (1986). “Seismic resistance of frames incorporating precast prestressed concrete beam shells.” PCI J., 31(4), 54–93.
PCI (Precast/Prestressed Concrete Institute). (2010). PCI design handbook,
7th Ed., Chicago.
Perez, F., Pessiki, S., and Sause, R. (2004a). “Lateral load behavior of
unbonded post-tensioned precast concrete walls with vertical joints.”
PCI J., 49(2), 48–64.
03118001-16
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
Perez, F., Pessiki, S., and Sause, R. (2004b). “Seismic design of unbonded
post-tensioned precast concrete walls with vertical joint connectors.”
PCI J., 49(1), 58–79.
Perez, F. J., Pessiki, S., and Sause, R. (2013). “Experimental lateral load
response of unbonded post-tensioned precast concrete walls.” ACI
Struct. J., 110(6), 1045–1055.
Perez, F. J., Sause, R., and Pessiki, S. (2007). “Analytical and experimental
lateral load behavior of unbonded posttensioned precast concrete
walls.” J. Struct. Eng., 10.1061/(ASCE)0733-9445(2007)133:11(1531),
1531–1540.
Pincheira, J. A., Oliva, M. G., and Kusumo-Rahardjo, F. I. (1998). “Tests
on double tee flange connectors subjected to monotonic and cyclic
loading.” PCI J., 43(3), 82–96.
Priestley, M. (1991). “Overview of PRESSS research program.” PCI J.,
36(4), 50–57.
Priestley, M., Sritharan, S., Conley, J., and Pampanin, S. (1999). “Preliminary results and conclusions from the PRESSS five-story precast concrete test building.” PCI J., 44(6), 42–67.
Priestley, M., and Tao, J. (1993). “Seismic response of precast prestressed
concrete frames with partially debonded tendons.” PCI J., 38(1), 58–69.
Psycharis, I. N., and Mouzakis, H. P. (2012). “Shear resistance of pinned
connections of precast members to monotonic and cyclic loading.”
Eng. Struct., 41, 413–427.
PTI (Post-Tensioning Institute). (2006). Post-tensioning manual, 6th Ed.,
Farmington Hills, MI.
Rahman, A., and Restrepo, J. (2000). “Earthquake resistant precast concrete buildings: Seismic performance of cantilever walls prestressed
using unbonded tendons.” Rep. No. 2000-5, Univ. of Canterbury,
Christchurch, New Zealand.
Rahman, M. A., and Sritharan, S. (2015). “Seismic response of precast,
posttensioned concrete jointed wall systems designed for low- to midrise buildings using the direct displacement-based approach.” PCI J.,
60(2), 38–56.
Ren, R., and Naito, C. (2013). “Precast concrete diaphragm connector performance database.” J. Struct. Eng., 10.1061/(ASCE)ST.1943-541X
.0000598, 15–27.
Restrepo, J. I. (2003). “Self-centering precast post-tensioned cantilever
walls: Theory and experimental work.” Structures Congress, ASCE,
Seattle.
Restrepo, J. I., Park, R., and Buchanan, A. H. (1995). “Tests on connections
of earthquake resisting precast reinforced concrete perimeter frames of
buildings.” PCI J., 40(4), 44–61.
Restrepo, J. I., and Rahman, A. (2007). “Seismic performance of selfcentering structural walls incorporating energy dissipators.” J. Struct.
Eng., 10.1061/(ASCE)0733-9445(2007)133:11(1560), 1560–1570.
Restrepo, J. I., Tobolski, M. J., and Matsumoto, E. E. (2011). “Development of a precast bent cap system for seismic regions.” NCHRP Rep.
681, Transportation Research Board, Washington, DC, 106.
Rodriguez, M., Restrepo, J. I., and Blandón, J. J. (2007). “Seismic design
forces of rigid floor diaphragms in precast concrete building structures.” J. Struct. Eng., 10.1061/(ASCE)0733-9445(2007)133:11(1604),
1604–1615.
Rodriguez, M., Restrepo, J. I., and Carr, A. J. (2002). “Earthquake induced
floor horizontal accelerations in buildings.” Earthquake Eng. Struct.
Dyn., 31(3), 693–718.
Saatcioglu, M., et al. (2001). “The August 17, 1999, Kocaeli (Turkey)
earthquake—Damage to structures.” Can. J. Civ. Eng., 28(4),
715–737.
Sakai, J., Unjoh, S., and Hoshikuma, J. (2009). “Development of seismic
design method for precast segmental concrete bridge column.” Technical Rep. MCEER-09-0012, Multidisciplinary Center for Earthquake
Engineering Research, Buffalo, NY.
Schoettler, M. J., Belleri, A., Zhang, D., Restrepo, J., and Fleischman, R. B.
(2009). “Preliminary results of the shake-table testing for development
of a diaphragm seismic design methodology.” PCI J., 54(1), 100–124.
Seeber, K. (2014). “The future use of PC in China.” Technical Presentation, Southeast Univ., Nanjing, P.R. China.
Seifi, P., Henry, R. S., and Ingham, J. M. (2016). “Panel connection details
in existing New Zealand precast concrete buildings.” Bull. N. Z. Soc.
Earthquake Eng., 49(2), 190–199.
© ASCE
Seifi, P., Henry, R. S., and Ingham, J. M. (2017). “In-plane testing of
precast concrete walls with grouted connections.” 16th World Conf.
in Earthquake Engineering, International Association of Earthquake
Engineering, Tokyo.
Seo, C., and Sause, R. (2005). “Ductility demands on self-centering
systems under earthquake loading.” ACI Struct. J., 102(2), 275–285.
Shaikh, A. F., and Feile, E. P. (2004). “Load testing of a precast concrete
double-tee flange connector.” PCI J., 49(3), 85–94.
Shim, C. S., Chung, C. S., and Kim, H. H. (2008). “Experimental evaluation of seismic performance of precast segmental bridge piers with a
circular solid section.” Eng. Struct., 30(12), 3782–3792.
Shutt, C. A. (1997). “Hybrid precast frame meets seismic challenges.”
Ascent, Spring, 14–19.
Sideris, P., Aref, A. J., and Filiatrault, A. (2014a). “Effects of anchorage
hardware on the cyclic tensile response of unbonded monostrands.”
PCI J., 59(3), 60–77.
Sideris, P., Aref, A. J., and Filiatrault, A. (2014b). “Large-scale seismic
testing of a hybrid sliding-rocking posttensioned segmental bridge
system.” J. Struct. Eng., 10.1061/(ASCE)ST.1943-541X.0000961,
04014025.
Sideris, P., Aref, A. J., and Filiatrault, A. (2015). “Experimental seismic
performance of a hybrid sliding: Rocking bridge for various specimen
configurations and seismic loading conditions.” J. Bridge Eng., 10.1061
/(ASCE)BE.1943-5592.0000742, 04015009.
Smith, B., and Kurama, Y. (2014). “Seismic design guidelines for solid and
perforated hybrid precast concrete shear walls.” PCI J., 59(3), 43–59.
Smith, B., Kurama, Y., and McGinnis, M. (2011). “Design and measured
behavior of a hybrid precast concrete wall specimen for seismic
regions.” J. Struct. Eng., 10.1061/(ASCE)ST.1943-541X.0000327,
1052–1062.
Smith, B., Kurama, Y., and McGinnis, M. (2013). “Behavior of precast
concrete shear walls for seismic regions: Comparison of hybrid and emulative specimens.” J. Struct. Eng., 10.1061/(ASCE)ST.1943-541X
.0000755, 1917–1927.
Smith, B., Kurama, Y., and McGinnis, M. (2015). “Perforated hybrid precast shear walls for seismic regions.” ACI Struct. J., 112(3), 359–370.
Sritharan, S. (2002). “Performance of four jointed precast frame systems
under simulated seismic loading.” 7th U.S. National Conf. on Earthquake Engineering, Earthquake Engineering Research Institute,
Oakland, CA.
Sritharan, S., Aaleti, S., Henry, R. S., Liu, K. Y., and Tsai, K. C. (2015).
“Precast concrete wall with end columns (PreWEC) for earthquake
resistant design.” Earthquake Eng. Struct. Dyn., 44(12), 2075–2092.
Sritharan, S., Priestley, M. J. N., and Seible, F. (2001). “Seismic design and
experimental verification of concrete multiple column bridge bents.”
ACI Struct. J., 98(3):335–346.
Standards New Zealand. (2004). “Structural design actions. Part 5:
Earthquake actions—New Zealand.” NZS 1170.5, Wellington,
New Zealand.
Standards New Zealand. (2006). “Concrete structures standard:
Amendment 2.” NZS 3101, Wellington, New Zealand.
Stone, W. C., Cheok, G. S., and Stanton, J. F. (1995). “Performance of
hybrid moment-resisting precast beam-column concrete connections
subjected to cyclic loading.” ACI Struct. J., 92(2), 229–249.
Taira, Y., Sakai, J., and Hoshikuma, J. (2009). “A study on restorable precast and prestressed hybrid piers.” Technical Rep. MCEER-09-0012,
Multidisciplinary Center for Earthquake Engineering Research,
Buffalo, NY.
Tazarv, M., and Saiidi, M. (2015). “UHPC-filled duct connections for accelerated bridge construction of RC columns in high seismic zones.”
Eng. Struct., 99, 413–422.
Thonstad, T., Mantawy, I. M., Stanton, J. F., Eberhard, M. O., and Sanders,
D. H. (2016). “Shaking table performance of a new bridge system with
pretensioned rocking columns.” J. Bridge Eng., 10.1061/(ASCE)BE
.1943-5592.0000867, 04015079.
Tobolski, M. J., and Restrepo, J. I. (2008). “Development of rocking column systems.” 6th National Seismic Conf. on Bridges and Highways,
M. Keever and L. Mesa, eds., Univ. at Buffalo, Buffalo, NY.
03118001-17
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Downloaded from ascelibrary.org by University of Notre Dame on 01/17/18. Copyright ASCE. For personal use only; all rights reserved.
Toniolo, G., and Colombo, A. (2012). “Precast concrete structures: The
lessons learned from the L’Aquila earthquake.” Struct. Concr., 13(2),
73–83.
Torres Matos, M. Y., and Rodriguez, M. (2014). “Comportamiento Sísmico
de Cimentación Tipo Candelero para Puentes Con Columnas Prefabricadas de Concreto Reforzado.” Revista de Ingenieria Sísmica, 90,
55–87 (in Spanish).
Trono, W., Jen, G., Panagiotou, M., Schoettler, M., and Ostertag, C. P. (2014).
“Seismic response of a damage-resistant recentering posttensionedHYFRC bridge column.” J. Bridge Eng., 10.1061/(ASCE)BE.1943-5592
.0000692, 04014096.
Twigden, K. M., Henry, R. S., and Sritharan, S. (2017). “Cyclic testing
of unbonded post-tensioned concrete wall systems with and without
supplemental damping.” Eng. Struct., 140, 406–420.
UBC (Uniform Building Code). (1997). “International conference of building officials.” UBC-97, Whittier, CA.
Vander Werff, J., Snyder, R., Sritharan, S., and Holombo, J. (2015). “A
cost-effective integral bridge system with precast concrete I-girders
for seismic application.” PCI J., 60(5), 76–95.
Veletzos, M. J., and Restrepo, J. (2009). “Influence of vertical earthquake
motion and pre-earthquake stress on joint response of precast concrete
segmental bridges.” PCI J., 54(3), 99–128.
Veletzos, M. J., and Restrepo, J. I. (2010). “Modeling of jointed connections in segmental bridges.” J. Bridge Eng., 10.1061/(ASCE)BE.19435592.0000112, 139–147.
Veletzos, M. J., and Restrepo, J. I. (2011). “Development of seismic design
guidelines for segmental construction.” Rep. No. SSRP-10, Univ. of
California, San Diego, 357–384.
Venuti, W. J. (1970). “Diaphragm shear connectors between flanges of
prestressed concrete T-beams.” PCI J., 15(1), 67–78.
Walsh, K., Draginis, R., Estes, R., and Kurama, Y. (2015). “Effects of
anchor wedge dimensional parameters on post-tensioning strand performance.” PCI J., 60(3), 63–83.
Walsh, K., Henry, R., Simkin, G., Brooke, N., Davidson, B., and Ingham, J.
(2016). “Testing of reinforced concrete frames extracted from a building
damaged during the Canterbury earthquakes.” ACI Struct. J., 113(2),
349–362.
Walsh, K., and Kurama, Y. (2010). “Behavior of unbonded post-tensioning
monostrand anchorage systems under monotonic tensile loading.”
PCI J., 55(1), 97–117.
Walsh, K., and Kurama, Y. (2012). “Effects of loading conditions on
the behavior of unbonded post-tensioning strand-anchorage systems.”
PCI J., 57(1), 76–96.
Wan, G., Fleischman, R. B., and Zhang, D. (2012). “Effect of spandrel
beam to double tee connection characteristic of flexure-controlled
© ASCE
View publication stats
precast diaphragms.” J. Struct. Eng., 10.1061/(ASCE)ST.1943-541X
.0000426, 247–257.
Wang, J. C., Ou, Y. C., Chang, K. C., and Lee, G. C. (2008). “Large-scale
seismic tests of tall concrete bridge columns with precast segmental
construction.” Earthquake Eng. Struct. Dyn., 37(12), 1449–1465.
Watkins, J., Sritharan, S., and Henry, R. (2014). “An experimental investigation of a wall-to-floor connector for self-centering walls.” 10th
U.S. National Conf. on Earthquake Engineering, Earthquake Engineering Research Institute, Oakland, CA.
Weldon, B., and Kurama, Y. (2007). “Nonlinear behavior of precast concrete coupling beams.” J. Struct. Eng., 10.1061/(ASCE)0733-9445
(2007)133:11(1571), 1571–1581.
Weldon, B., and Kurama, Y. (2010). “Experimental evaluation of posttensioned precast concrete coupling beams.” J. Struct. Eng., 10.1061
/(ASCE)ST.1943-541X.0000212, 1066–1077.
Weldon, B., and Kurama, Y. (2012). “Analytical modeling and design validation of posttensioned precast concrete coupling beams for seismic
regions.” J. Struct. Eng., 10.1061/(ASCE)ST.1943-541X.0000415,
224–234.
White, S., and Palermo, A. (2016). “Quasi-static testing of posttensioned
nonemulative column-footing connections for bridge piers.” J. Bridge
Eng., 10.1061/(ASCE)BE.1943-5592.0000872, 04016025.
Wood, S. L., Stanton, J. F., and Hawkins, N. M. (1995). “Performance
of precast parking garages during the 1994 Northridge earthquake.”
13th ASCE Structures Congress, ASCE, New York, 563–566.
Wood, S. L., Stanton, J. F., and Hawkins, N. M. (2000). “New seismic
design provisions for diaphragms in precast concrete parking structures.” PCI J., 45(1), 50–65.
Yamashita, R., and Sanders, D. H. (2009). “Seismic performance of precast
unbonded prestressed concrete columns.” ACI Struct. J., 106(6),
821–830.
Zhang, D., and Fleischman, R. B. (2016). “Establishment of performancebased seismic design factors for precast concrete floor diaphragms.”
Earthquake Eng. Struct. Dyn., 45(5), 675–698.
Zhang, D., Fleischman, R. B., Naito, C., and Ren, R. (2011). “Experimental
evaluation of pretopped precast diaphragm critical flexure joint under
seismic demands.” J. Struct. Eng., 10.1061/(ASCE)ST.1943-541X
.0000352, 1063–1074.
Zheng, W. (2001). “Analytical method for assessment of seismic shear
capacity demand for untopped precast double-tee diaphragms joined
by mechanical connectors.” Ph.D. dissertation, Univ. of WisconsinMadison, Madison, WI.
Zheng, W., and Oliva, M. G. (2005). “A practical method to estimate elastic
deformation of precast pretopped double tee diaphragms.” PCI J.,
50(2), 44–55.
03118001-18
J. Struct. Eng., 2018, 144(4): 03118001
J. Struct. Eng.
Download