Sei sulla pagina 1di 25

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

net/publication/309965802

Development and Seismic Behavior of Precast Concrete Beam-to-Column


Connections

Article  in  Journal of Earthquake Engineering · November 2016


DOI: 10.1080/13632469.2016.1217807

CITATIONS READS

18 4,935

4 authors, including:

Cheng Jiang Hanbin Ge


The Hong Kong Polytechnic University Meijo University
36 PUBLICATIONS   396 CITATIONS    167 PUBLICATIONS   1,615 CITATIONS   

SEE PROFILE SEE PROFILE

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

Experimental Study and Analytical Modeling of FRP Rehabilitated Concrete Members View project

Coupling effect of buckling and ductile fracture for steel structures View project

All content following this page was uploaded by Cheng Jiang on 20 October 2017.

The user has requested enhancement of the downloaded file.


Journal of Earthquake Engineering

ISSN: 1363-2469 (Print) 1559-808X (Online) Journal homepage: http://www.tandfonline.com/loi/ueqe20

Development and Seismic Behavior of Precast


Concrete Beam-to-Column Connections

Dongzhi Guan, Cheng Jiang, Zhengxing Guo & Hanbin Ge

To cite this article: Dongzhi Guan, Cheng Jiang, Zhengxing Guo & Hanbin Ge (2016):
Development and Seismic Behavior of Precast Concrete Beam-to-Column Connections, Journal
of Earthquake Engineering, DOI: 10.1080/13632469.2016.1217807

To link to this article: http://dx.doi.org/10.1080/13632469.2016.1217807

Accepted author version posted online: 11


Nov 2016.
Published online: 11 Nov 2016.

Submit your article to this journal

Article views: 37

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=ueqe20

Download by: [City University of Hong Kong Library] Date: 20 November 2016, At: 02:03
Journal of Earthquake Engineering, 00:1–23, 2016
Copyright © Taylor & Francis Group, LLC
ISSN: 1363-2469 print/1559-808X online
DOI: 10.1080/13632469.2016.1217807

Development and Seismic Behavior of Precast


Concrete Beam-to-Column Connections

DONGZHI GUAN1,2,3, CHENG JIANG4, ZHENGXING GUO1,2,


and HANBIN GE3
1
School of Civil Engineering, Southeast University, Nanjing, China
2
Key Laboratory of Concrete and Pre-Stressed Concrete Structures of the
Ministry of Education, Nanjing, China
3
Department of Civil Engineering, Meijo University, Nagoya, Japan
4
Department of Civil and Architectural Engineering, City University of Hong
Kong, Hong Kong SAR, China

A new precast concrete beam-to-column connection for moment-resisting frames was developed in this
study. Both longitudinal bar anchoring and lap splicing were used to achieve beam reinforcement
continuity. Three full-scale beam-to-column connections, including a reference monolithic specimen,
were investigated under reversal cyclic loading. The difference between the two precast specimens was
the consideration of additional lap-splicing bars in the calculation of moment-resisting strength. Seismic
performance was evaluated based on hysteretic behavior, strength, ductility, stiffness, and energy
dissipation. The plastic hinge length of the specimens is also discussed. The results show that the
proposed precast system performs satisfactorily under reversal cyclic loading compared with the
monolithic specimen, and the additional lap-splicing bars can be included in the strength calculation
using the plane cross-section assumption. Furthermore, the plastic hinge length of the proposed precast
beam-to-column connection can be estimated using the models for monolithic specimens.

Keywords Precast Concrete; Beam-to-Column Connection; Cyclic Loading; Seismic Performance;


Plastic Hinge Length

1. Introduction
Compared with conventional cast-in-place concrete structures, precast concrete systems are
advantageous in terms of product quality, cost efficiency, and speed of construction [Crisafulli
and Restrepo, 2003; Korkmaz and Tankut, 2005; Shariatmadar and Zamani, 2014]. To save
natural resources and reduce environmental pollution, precast concrete structures are also
considered as environmentally friendly and green constructions [Guo, 2014; Yee, 2001]. In
many countries including the United States, Japan, and China, precast concrete systems have
been increasingly implemented for both residential and industrial buildings. Regarding
structural components, beams and columns are more suitable for standardization, finalization,
and modularization, which lead to precast concrete frames being a better option for the precast
concrete industry. However, beam-to-column connections act as vital parts in precast concrete
frames [Choi et al., 2013; Le-Trung et al., 2013], not only affecting the overall structural
behavior of precast concrete frames but also planning for the manufacture and installation of

Received 1 April 2016; accepted 21 July 2016.


Address correspondence to Zhengxing Guo, School of Civil Engineering, Southeast University, No. 2,
Sipailou, Nanjing 210096, China. E-mail: guozx195608@126.com
Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/ueqe.

1
2 D. Guan et al.

precast concrete members. Therefore, great efforts have been made to research the various
types of beam-to-column connections including emulative, pretensioned, welded, bolted,
doweled, and hybrid connections [ACI-ITG 1, 2003; Ericson, 2010; fib Bulletin 43, 2008;
Li et al., 2009].
Among the existing precast concrete beam-to-column connections, emulative precast
systems are commonly adopted in the precast concrete industry [Park, 1995] because they
are easy to be designed and detailed with nearly the same seismic performance as
conventional monolithic structures following the principle of emulating monolithic rein-
forced concrete (RC) construction. Therefore, research on emulative precast connections
remains one of the promising areas that require many efforts.
Park and Bull [1986] experimentally investigated composite beam-exterior column
subassemblies, incorporating precast prestressed concrete beam shells. It was concluded
that the units would behave satisfactorily in ductile seismic-resisting frames if they were
designed for seismic loading. Restrepo et al. [1995] tested two full-scaled cruciform
specimens of beam-to-column connections under quasi-static cyclic reversed loading. In
one specimen, the bottom bars of composite beams were anchored inside the joint core
directly. The results showed that the connection detail could be successfully designed and
constructed to emulate cast-in-place construction. Investigations on the seismic behavior
and moment rotation behavior of unbonded post-tensioned precast concrete beam-to-
column connections were carried out at the Center for Advanced Technology for Large
Structural Systems at Lehigh University [El-Sheikh et al., 1999, 2000]. Different analy-
tical models were developed for unbonded post-tensioned precast frames. It was con-
cluded that the behavior of unbonded post-tensioned precast frames was more than
adequate for severe earthquake loading.
Alcocer et al. [2002] utilized unidirectional and bidirectional cyclic loading to test two
different precast beam-to-column connections in a precast concrete frame. Conventional mild
reinforcing bars and prestressing strands were used to achieve beam continuity in the tested
connections. Ali and Parastesh [2003] proposed and studied a simple moment-resisting precast
concrete beam-to-column connection for highly seismic regions. In the proposed connection,
beam reinforcement continuity was created using a combination of lap splicing and end
anchorage of bars. The results showed that the proposed precast connection was capable of
matching or exceeding the performance of the monolithic connection. Ertas et al. [2006]
experimentally studied precast beam-to-column connections, including the cast-in-place in
column connection, the cast-in-place in beam connection, and the composite connection. The
results revealed that the three types of precast connections were suitable for high-seismic
regions. Xue and Yang [2010] exhibited experiments on precast concrete connections at
different locations in a moment-resisting frame and concluded that the precast connections
resulted in a strong-column, weak-beam failure model with plastic hinges at the beam ends.
Hosoya et al. [2012] tested a new precast concrete system in which beam-to-column
joints were manufactured as precast concrete in Japan. Vertical holes were made in the
prefabricated beam-to-column joints for reinforcing bars of the lower precast column to
pass through, and reinforcing bars of the precast beams were connected by sleeves. Cai
et al. [2012] investigated beam-to-column connections of so-called SCOPE systems.
U-shaped conventional deformed bars were adopted to overlap the strands, which were
used as the bottom longitudinal bars of the beams. Chen et al. [2012] presented an
experimental and analytical investigation on large-scale interior precast beam-to-column
joints with slabs. The precast beam reinforcement on the bottom was formed and installed
with an anchor head. The test results indicated that the performance of the joint was
satisfactory.
Development and Seismic Behavior of Precast Beam-column Connections 3

Maya et al. [2013] suggested the use of ultra high-performance fiber-reinforced con-
cretes in precast beam-to-column connections to reduce the lengths of reinforcement splice.
A seismic test was conducted, and the results showed that improved detailing should be
considered to increase the energy dissipation capacity. Im et al. [2013] conducted reversal
cyclic loading test of six cruciform precast beam-to-column connections. Straight reinforcing
bars were added into the beam-to-column joints and the beam core of U-shape shells to lap
splice the reinforcing bars of the beams. The results revealed that the deformation capacities
of these precast beam-to-column connections were comparable to those of monolithic
connections, whereas the stiffness and energy dissipation were unsatisfactory.
Ha et al. [2014] developed a new concrete beam-to-column joint comprising precast
concrete (PC) beams with U-shaped strands, PC columns, PC slabs, and topping mono-
lithic concrete. Three interior and three exterior semi-PC joint specimens were tested
under the lateral cyclic load. The test results showed that the proposed structural system
with transverse reinforcements at the joint performed satisfactorily for moderate seismic
regions. Parastesh et al. [2014] proposed a new ductile moment-resisting connection in
precast concrete frames. Columns were prefabricated with a free space in the connection
zone, and precast beams were manufactured with U-shaped cross sections at the ends.
Cast-in-situ concrete and lap-splicing bars were used to ensure the integrity of precast
elements. The test results showed that the proposed system could provide an adequate
seismic resistance capacity.
It can be observed that most emulative precast joints are wet connections using cast-
in-place techniques on site [Pampanin et al., 2001]. To achieve reinforcement continuity,
the bottom longitudinal bars of precast beams either extrude and anchor into joint cores or
are just lap-spliced by additional bars crossing joint cores. However, there are scarce
precast beam-to-column connections combining both aforementioned methods. For the
precast concrete connection with anchored longitudinal bars, the field application is
considered fairly inflexible because the relatively large number of anchored straight bars
increases the conflicting risk and reduces the speed of erection at sites in practice.
Regarding the means of lap splicing, the seismic performance of precast connections is
deemed inferior to monolithic ones, especially in high-seismic regions. It is due to the fact
that the regions for lap splicing are always the plastic hinge zones, and damage to the
concrete holding lap splicing bars would weaken the connecting capacity of lap splicing
quickly.
In existing precast concrete systems, there is no combination of anchored reinforcing
bars supplemented by hooked bars to connect precast components of moment frames. This
paper proposes a new precast concrete beam-to-column connection, adopting both long-
itudinal bar anchoring and lap splicing to achieve beam reinforcement continuity for the
purpose of easy construction and satisfactory seismic resistance as good as monolithic
connections. Although it has been adopted and constructed in a building that is more than
80 m tall in China, no experiments have been performed to test the seismic performance of
this connection under reversal cyclic loading. In this investigation, two precast beam-to-
column specimens and a reference monolithic specimen were tested to evaluate the
effectiveness of this precast connection.

2. Concept of Developed Connection


The new precast beam-to-column connection proposed in this paper is characterized by
both anchoring and lap splicing to achieve continuity of precast beam bottom reinforce-
ments. Figure 1 depicts the details of the precast components in the developed beam-to-
column connection. Prefabricated columns are manufactured with a height of one story,
4 D. Guan et al.

FIGURE 1 Connecting zone details of the precast components.

and grout sleeves are installed in the bottom of each precast column to connect the
longitudinal bars of the lower column. Precast beams are prefabricated with a hollow
U-shaped cross sections at the beam ends. The vertical side shells at the U-section zones
are intended to be 50-mm thick to fix the stirrups, and the thickness of the bottom shells at
the beam ends must be designed to hold the bottom longitudinal bars of the precast beams.
The top surfaces of the precast beams are roughened to increase the bond capability
between the precast concrete and the cast-in-place concrete. The surfaces of the U-shaped
shells at the precast beams’ ends are kept smooth for the ease of removing formworks
from the U-shaped hollows during the prefabrication of precast beams.
The arrangement of reinforcing bars in prefabricated beams is difficult to consider for
conventional emulative precast connections, which adopt anchoring to achieve continuity
of the bottom longitudinal bars in precast beams. This is because the longitudinal bars of a
precast beam must be kept away from those of precast columns and other precast beams.
Small errors of laying reinforcements would make the whole precast component useless.
In this proposed system, the only two bottom longitudinal bars of a precast beam are
allowed to extend out from the bottom shells and anchor into a joint core. Anchorage
plates are welded to the end of each extruding bar to enhance the anchoring capacity. In a
typical precast beam-to-column connection suggested, two bottom longitudinal bars of one
precast beam are placed at the corners of the bottom shells, whereas those of the other
precast beam are laid at the middle of the bottom shells to avoid obstruction. Usage of
fewer anchoring longitudinal bars greatly simplifies the prefabrication of elements and
erection of precast structures. To ensure the load-bearing capacity of the beams, reinfor-
cing bars of large diameters, such as 22 and 25 mm, are chosen. Considering the
discontinuity between precast elements, additional U-shaped bars can be added into the
connection zone. Open stirrups with strict 135° hooks are used in U-section zones to
simplify the placement of additional bars into the connection zones. The spacing of
stirrups in this precast system can be determined by the same confinement strategy as
conventional monolithic concrete structures.
Precast composite slabs are used in the proposed system, which consist of prefabri-
cated parts in the bottom and monolithic parts in the top. The site operation sequence of
each floor is as follows: (a) erect the columns and grout the column bars, (b) set up the
Development and Seismic Behavior of Precast Beam-column Connections 5

FIGURE 2 Details of the precast beam-to-column connection.

loose scaffolds for precast horizontal members, (c) set the precast beams and prefabricated
slabs in place, (d) add the loose bars and ties and fix the formwork around the joint, (e)
pour the joints and top parts of the slabs, and (f) cure the poured concrete and wait for it to
harden. Figure 2 shows the integrity of the developed precast beam-to-column connection.
The proposed precast connections can be designed easily using the same design strategy
as conventional monolithic connections. In practice, the bottom longitudinal bars of beams
in the proposed system can be considered to carry all design loads, and additional bars are
added to the connection to compensate for the discontinuity between precast elements by
experience. For economic reasons, the additional bars can also be included in the calculation
of flexural strength based on the plane cross-section assumption.

3. Experimental Investigation
Reversed cyclic loading tests of three full-scale cruciform specimens, including a reference
monolithic specimen, were conducted to evaluate the seismic performance of the proposed
beam-to-column connection.

3.1. Description of Specimens


Specimens were considered to represent an 82.6-m–tall building in the east of China,
which consisted of precast components of moment frames and monolithic shear walls. In
the prototype structure, the span/depth ratios of beams ranged from 8.16 to 14.0, the
reinforcement ratio of most beams was approximately 2%, and the typical column section
was 800 mm × 800 mm. The columns near the center of the prototype building were
designed to bear the axial load of approximately 0.5 fcAg, where fc was the compressive
strength of concrete for the design specified by the Chinese Code for design of concrete
structures [GB50010-2010, 2010], and Ag was the sectional area of the column. A strong-
column, weak-beam design philosophy was adopted to ensure beam sideway mechanism
under seismic loading. Moreover, shear resistance of the test specimens was checked to
avoid shear failure before flexural failure of the beam-to-column connection. All detailing
of the specimens was designed to satisfy the code [GB50010-2010, 2010].
Two precast beam-to-column connection specimens were designated as S2 and S3;
meanwhile, the reference monolithic specimen was marked as S1. Their dimensions and
6 D. Guan et al.

reinforcement quantities were designed, considering both the prototype structure and the
test condition. The heights of the columns above and below the beams were 1375 and
1000 mm, respectively, equal to the height of one story on the test instruments. The
column longitudinal reinforcements consisted of four D25 deformed bars in the corners
and eight D20 bars in the sides of the 550 mm × 550 mm cross section. In the zone from
600 mm above to 600 mm below the beams, the spacing of column stirrups was 100 mm,
and out of that zone, the interval was 200 mm. The beam had a length of 2000 mm and a
300 mm × 450 mm cross section. Within the range of 0–1000 mm from the column face,
the beam stirrups were spaced at 100-mm intervals to enhance the beam ends. Other beam
stirrups maintained 200 mm spacing. D10 and D8 rectangular hoops acted as stirrups in
the columns and beams, respectively.
As shown in Fig. 3a, the beam longitudinal reinforcements of the monolithic speci-
men remained continuous throughout the two beams, including five D20 deformed bars at
the top and three D20 bars at the bottom. Figure 3b depicts the reinforcement details of
precast concrete specimens. D25 bars were adopted as the beam bottom longitudinal
reinforcements of Specimen S2, whereas the D22 bars were of Specimen S3. Two D14
deformed bars were formed into U shapes as additional bars in the connection zones.
Specimen S2 was designed by an easy and practical method including only the two D25
bars in the calculation of flexural strength. Specimen S3 was designed by an economical
design strategy including both the bottom longitudinal bars and the additional bars in the
calculation of flexural strength based on the plane cross-section assumption, which was
the same as the strength of Specimen S1. The two precast specimens were designed to be
equal to the monolithic specimen in terms of moment-resisting capacity. However, in
Specimen S2, the additional bars were regarded as a strengthening measure and were not
considered in the calculation of moment-resisting capacity. Other reinforcement details of
precast specimens were the same as the monolithic specimen. The bottom shells in the
U-shaped hollow section were 65-mm thick to fix D25 or D22 bars and D8 stirrups. The
lengths of the U-shaped hollows in the precast concrete beams were 500 mm to guarantee
the anchorage length of additional bars.
Concrete was cast three times: the monolithic specimen and precast columns were
made first, and then precast beams were poured. After the assembly of precast compo-
nents, connection zones and composite layers on precast beams were filled with concrete
during the third time.

3.2. Material Properties


Concrete of Grade C40 was intended to be used in the test specimens, of which the cubic
compressive strength should achieve approximately 40 MPa. In terms of reinforcing bars,
Grade HRB 400 was adopted, which meant that the yielding strength of reinforcing bars
should be 400 MPa. To determine the real strength of the used concrete, control concrete
cubes were made with the same poured concrete and cured in the same environment as the
specimens. The cubes were tested to obtain the cubic compressive strength at the same
time the cyclic loading experiments were conducted. The cubic compressive strength (fcu)
of concrete in different elements is given in Table 1. The relationship fc′ = 0.78 fcu [Lu
et al., 2005] was used to evaluate the cylinder compressive strength of concrete (fc′). The
yielding (fy) and tensile (fu) strengths of the reinforcing bars were obtained by uniaxial
tension testing and are also presented in Table 1.
Development and Seismic Behavior of Precast Beam-column Connections 7

(a)

(b)

FIGURE 3 Configurations and reinforcement details of the monolithic and precast concrete
specimens. (a) Monolithic Specimen S1. (b) Precast Specimens S2 and S3 (Units: mm).

3.3. Test Setup


All specimens were tested under reversed cyclic loading at the Key Laboratory of Civil
Engineering and Disaster Prevention in Jiangsu Province, China. As shown in Fig. 4, speci-
mens were installed on a column base supported by the bottom roller. The column’s top was
8 D. Guan et al.

TABLE 1 Yield and ultimate strength of the reinforcing bars


Concrete fcu (MPa) Bar diameter (mm) D8 D10 D20 D22 D25
S1/precast columns 55.5 Area (mm2) 50.2 78.5 314.0 379.9 490.6
Precast beams 51.4 fy (MPa) 448 433 448 450 429
Connection zones 56.1 fu (MPa) 646 598 617 624 607
D: bar diameter in millimeters.

FIGURE 4 Details of the test setup.

connected to a 1500 kN actuator through a hinge. The distance from the bottom hinge to the
center of the actuator was 2900 mm, which was equal to the height of one story. Beam ends
were restricted vertically by roller supports so that lateral movements were allowed. The
boundary conditions were intended to simulate the real situation, in which the moments
approached zero at the mid-span points of columns and beams under seismic conditions.
Four hydraulic jacks were placed on the steel girders seated on the tops of the columns to
simulate gravity loads. Four bundles of prestressing strands were mounted before and behind
the column symmetrically, which were used to transmit axial forces applied by hydraulic jacks
to the column base. Thus, the hinge under the column base need not bear column axial
loadings. The P−Δ effect was not considered because the direction of hydraulic jacks
corresponded to the column throughout the test process.

3.4. Loading Sequence


Considering the high dead load transferred from the upper floors of high-rise residential
buildings, an axial force of 0.22 fc′Ag was applied onto the top of the column by the four
Development and Seismic Behavior of Precast Beam-column Connections 9

FIGURE 5 Cyclic loading history.

hydraulic jacks and was held constant throughout the test procedure. The applied axial
force equaled the design axial load of 0.5 fcAg, which was intended to simulate the interior
frames of the prototype building. Displacement-controlled lateral cyclic loading was
applied by the 1500 kN actuator according to the loading history, as shown in Fig. 5.
The drift angle was computed as the ratio of the actuator displacement to the distance from
the lateral loading point to the hinge under the column base. The first two drift levels of
0.034% and 0.069% were applied for one cycle as a pretest to check the whole test system.
After that, each specimen was gradually loaded to drift angles of 0.2%, 0.25%, 0.35%,
0.5%, 0.75%, 1%, 1.5%, 2%, 2.75%, 3.5%, and 4.25%, and each drift level was applied
for three cycles. If the peak load of a loading cycle decreased to less than 80% of the
maximum load during the test, the test specimen would be considered destroyed, and the
cyclic test would be terminated.

4. Experimental Results and Discussions

4.1. Failure Mode


Figure 6a, 6b, and 6c compare the final failure modes of the three specimens at the end of
the cyclic loading test. In monolithic Specimen S1, no cracks appeared during the pretest.
After the first cycle of 0.2% drift level, four fine flexural cracks were observed on a beam
when the bottom fibers of the beam were in tension, and the four cracks were located at
distances of 5, 10, 35, and 62 cm from the face of the column. Lateral compressive cracks
occurred at the bottom of a beam when the loading drift reached 2.75%, indicating that the
bottom concrete of the beam began to crush. During the 3.5% drift loading, many concrete
blocks dropped from the bottoms of beam ends near the column. At the first cycle of
4.25% drift loading, the bottom longitudinal bars of beams could be observed to have
bended in compression, and the applied load became less than 80% of the maximum load.
As a result, the test was stopped.
In precast Specimen S2, a pretest also did not cause any cracks in the specimen. When
the first cycle of 0.2% was finished, two flexural cracks occurred in the beam at distances
of approximately 20 and 50 cm from the column face. The location 50 cm from the
column face was the place where cast-in-place and precast concrete interfaced in the
U-shaped region, and the crack that occurred there was wider and longer than the other
one. As the loading drift increased, there were some main inclined cracks propagating
10 D. Guan et al.

(a)

(b) (c)

FIGURE 6 Failure modes of the specimens at the end of the test: (a) S1; (b) S2; and (c) S3.

gradually in the connecting region of the beams. During the 2.75% drift loading, some
small concrete blocks dropped from the middle of the section height at the beam ends.
When the loading drift reached 4.25%, quite a significant amount of concrete spalled at
the tops and bottoms of the beam ends near the column. At the third cycle of 4.25% drift
loading, the applied load dropped significantly, so the loading of Specimen S2 was
terminated.
The overall experimental behavior of Specimen S3 was similar to that of Specimen
S2. To find the first flexural crack, another drift loading of 0.1% was conducted after the
pretest loading. As a result of the 0.1% drift loading, only one crack appeared at a distance
of approximately 50 cm from the column face, where cast-in-place concrete encountered
precast concrete. During the 0.75% drift loading, a 20-cm-long lateral crack occurred at
the interface between the precast beam and the monolithic composite layer. The bottom
concrete of a beam near the column began to crush at a drift level of 3.5%, and the bottom
concrete in the region 0–50 cm away from the column face was totally damaged at 4.25%
drift loading. The test was ended upon the dramatic decrease of applied load at the third
cycle of 4.24% drift.
Generally, the three specimens failed with similar modes, developing plastic hinges at the
beam ends near the column face without any damage at the beam-to-column joints. This was
due to the strong-column, weak-beam design philosophy. Most concrete cracks were con-
centrated in the plastic regions of beams, and severe concrete spalling occurred at the bottoms
of beam ends near the column. The crushing of concrete at the tops of beam ends seemed less
severe than that at the bottom because there were more reinforcing bars at the tops of the
beams than at the bottoms. In the precast specimens, the first crack appeared near the
interfacial section of the cast-in-place and precast concrete, demonstrating that the interface
Development and Seismic Behavior of Precast Beam-column Connections 11

constituted the weakest section to resist flexural cracking. The concrete at the tops of beams in
precast Specimen S2 became more severely damaged than that in Specimen S3 because the
maximum load bear by Specimen S2 was higher than that by S3.

4.2. Hysteretic Behavior


Figure 7a, 7b, and 7c illustrate the moment–drift angle relationships of the monolithic and
precast specimens. The moment was evaluated as the product of the applied load and the
distance from the loading point to the column base hinge. At the early stage of loading, the
test specimens exhibited approximately elastic behavior and dissipated little energy. They
all began to yield at the drift angle of approximately 1%. After that, the hysteresis loops
gradually became fatter with the growth of drift loading, which indicated the increasing
energy-dissipating capacity. The peak load of the first cycle remained higher than those of
the following cycles at the same loading drift, representing strength degradation. Slight
pinching was observed before the specimens approached total destruction because of the
developing cracks and spalling of concrete.
In monolithic Specimen S1, the maximum strength was reached at 2.75% drift loading,
and greater pinching occurred after 3.5% drift loading compared with the precast specimens.
Only one cycle of 4.25% drift loading was applied before Specimen S1 was damaged
totally. Precast specimens also achieved the maximum strength at 3.5% drift loading and
bear three cycles of 4.25% drift loading, showing better hysteretic behavior.

(a)

(b) (c)

FIGURE 7 Hysteresis moment–drift angle relationships of the specimens: (a) S1; (b) S2;
and (c) S3.
12 D. Guan et al.

FIGURE 8 Moment–drift angle envelopes of the specimens.

Furthermore, the moment–drift angle envelopes are presented in Fig. 8. It can be


observed that there were obvious yielding stages of the three test specimens. The precast
specimens achieved higher strength than monolithic Specimen S1 after yielding and less
reduction in strength at the ends of loadings.

4.3. Flexural Strength and Strength Degradation


Because the test loading was controlled only by the drift angle, the yielding points
were defined by adopting the principle of equivalent elastoplastic energy. As shown in
Fig. 9, a bilinear curve is utilized to idealize the moment–drift angle envelope curve
with zero-post-yield stiffness. The area enclosed by the idealized curve should be equal
to that by the envelope curve, and the point on the envelope curve corresponding to
the turning point on the bilinear curve is defined as the yielding point. The yielding
and maximum strength are listed in Table 2. Although the three specimens were
designed to carry the same load using the plane cross-section assumption, the strength

FIGURE 9 Principle of equivalent elastoplastic energy.


Development and Seismic Behavior of Precast Beam-column Connections 13

TABLE 2 Moment-resisting capacities


Yielding Maximum Ratio of maximum to
Specimen Direction strength (kN) strength (kN) yielding strength Average
S1 Positive 483.00 518.78 1.07 1.08
Negative −460.72 −500.63 1.09
S2 Positive 550.80 593.46 1.08 1.08
Negative −515.74 −560.19 1.09
S3 Positive 506.83 545.72 1.08 1.07
Negative −496.25 −527.74 1.06

of precast specimens was found to be higher than that of the monolithic specimen. For
Specimen S2, additional bars were not included in the calculation of load-bearing
capacity, so the reinforcement ratio of Specimen S2 ranked first, which led to the
highest flexural strength. The strength of Specimen S3 became slightly higher than that
of Specimen S1, demonstrating that the calculation method was effective. The ratio of
maximum to yielding strengths was quite close, indicating similar safety assurance in
strength.
Figure 10a depicts the definition of strength degradation that occurs in the second and
third cycles of each drift loading [ATC, 2005]. The strength ratio αi is determined as the
average ratio of the peak load of the ith cycle to that of the first cycle in both positive and

(a)

(b) (c)

FIGURE 10 Cyclic strength degradation of the specimens. (a) Definition of the strength
ratio. (b) Strength ratio in the second load cycle. (c) Strength ratio in the third load cycle.
14 D. Guan et al.

negative directions at every drift angle. α2 and α3 are presented in Fig. 10b and c, respectively.
In the early stage of the test, the cyclic strength degradation of Specimens S2 and S3 seemed
slightly greater than that of monolithic Specimen S1 because of the weaker interfacial section
of cast-in-place and precast concrete. After the drift angle of 2.75%, α2 and α3 of Specimen S1
became smaller compared with Specimens S2 and S3, which represented better capacities of
precast specimens to maintain resisting strengths in the post-elastic range. α2 and α3 of precast
Specimens S2 and S3 dropped by an average of 0.93% and 2.9% for every 1% improvement in
the drift angle, respectively. As specified by ACI 374.1-05 [2005], α3 should not be less than
0.75 at a story drift ratio of 3.5%. α3 of Specimens S2 and S3 at a 3.5% drift angle was 0.92 and
0.89, respectively, which are greater than the acceptance criteria.

4.4. Ductility and Stiffness Degradation


Ductility is a vital factor in the seismic behavior of a structure or a component to avoid
brittle failure, describing the capacity to undergo inelastic deformation without significant
strength loss. The drift ductility is defined as the ratio of ultimate drift angle δu to yielding
drift angle δy. The yielding drift angle δy corresponds to the yielding point obtained by the
aforementioned equal energy principle. The ultimate drift angle δu is determined by the
failure point at which the strength decreases to 20% of the maximum load capacity.
Table 3 summarizes the yielding and ultimate drift angle of the specimens in two
directions as well as the ductility. The average ductilities of Specimens S1, S2, and S3
were 3.82, 3.97 and 4.07, respectively, which revealed the ductile behavior of test speci-
mens under reversal cyclic loading. The ductilities in the negative direction were margin-
ally less than those in the positive direction, with the slightly greater yielding drift angle in
the negative direction. In terms of ductilities, precast specimens showed better ductile
behavior than the monolithic specimen.
The stiffness degradation is another important parameter to assess a structure’s overall
response because it reflects the cumulative damage of a structure caused by seismic loads.
The peak-to-peak stiffness of each drift loading cycle is utilized to evaluate the stiffness
degradation. The peak-to-peak stiffness is defined as the slope of the secant line connecting
the peak response points in the positive and negative directions during a drift loading cycle
[Sucuoglu, 1995]. Figure 11 illustrates the secant stiffness calculated from the hysteresis
curves of the test specimens. Because some friction between the supporting plates and
rotating plate of the column base occurred during the rotation at drift angles from approxi-
mately 0.2% to 1%, the stiffness increased slightly in the drift angle range of 0.2–1%.
However, some conclusions could still be drawn from the degradation curves. The initial
stiffness of the monolithic specimen was greater than that of the precast specimens because
of predefined gaps between prefabricated components. Because the additional bars in precast

TABLE 3 Yielding and ultimate drift angle, ductilities


Specimen Direction Yielding drift angle (%) Maximum strength (%) δu/δy Average
S1 Positive 1.00 4.07 4.08 3.82
Negative −1.12 −4.00 3.57
S2 Positive 0.99 4.25 4.29 3.97
Negative −1.16 −4.25 3.65
S3 Positive 0.99 4.25 4.29 4.07
Negative −1.11 −4.25 3.85
Development and Seismic Behavior of Precast Beam-column Connections 15

FIGURE 11 Stiffness degradation of the specimens.

FIGURE 12 Normalized stiffness degradation of the specimens.

connections were located closer to the neutral axis of the beam section than the bottom
longitudinal bars, their yielding was delayed to a higher drift angle. Therefore, the stiffness
of Specimens S2 and S3 became greater than that of Specimen S1 after yielding.
To assess the ratio of stiffness degradation, peak-to-peak stiffness was divided by the
initial secant stiffness and normalized as a non-dimensional value (knorm). Figure 12
reveals the normalized stiffness of the test specimens. It could be observed that the
stiffness of Specimens S2 and S3 decreased slower than that of Specimen S1, with the
help of additional bars. The stiffness degradation of Specimen S3 was slightly more severe
than that of Specimen S2 as a result of the relatively lower bottom reinforcement ratio in
Specimen S3.
16 D. Guan et al.

4.5. Energy Dissipation


A beam-to-column connection with good energy dissipation capacity is considered to absorb a
sufficient amount of energy input by seismic excitation. It helps protect other structural
elements by reducing the transmitted energy and causes the whole structural system to perform
satisfactorily in the inelastic range. The cumulative dissipated energy and equivalent viscous
damping ratio are the main indexes to evaluate the energy dissipation capacity of a structure or
a component. The amount of dissipated energy in a given cycle is represented by the area
enclosed by the hysteretic loop in that cycle. The cumulative energy dissipated at a particular
drift angle is defined as the sum of the energy dissipated to that cycle. The equivalent viscous
damping ratio (ζeq) is expressed as the ratio of energy dissipated in a cycle to the strain energy
of an equivalent linear system multiplied by the constant 2П [Chopra, 1995]. Figures 13 and

FIGURE 13 Cumulative energy dissipation curves of the specimens.

FIGURE 14 Equivalent viscous damping ratio curves of the specimens.


Development and Seismic Behavior of Precast Beam-column Connections 17

14 show the cumulative dissipated energy and equivalent viscous damping ratio versus the
number of drift loading cycles of the specimens, respectively. The results of the pretest were
not included for the very little energy dissipated in the pretest stage. In Fig. 13, it is observed
that the three specimens dissipated almost the same amount of energy before the 19th loading
cycle. After the 19th loading cycle, the cumulative energy absorbed by Specimen S2 increased
faster than that dissipated by the other specimens because more reinforcing bars yielded, and
more concrete spalled in Specimen S2. Although the variation of the dissipated energy of
Specimens S1 and S3 seemed negligible, Specimen S3 dissipated more overall energy than
Specimen S1 because it bear two more loading cycles. In terms of the equivalent viscous
damping ratio, discriminative conclusions could be drawn. Before the 13th loading cycle, the
equivalent viscous damping ratio of the test specimens was approximated because the speci-
mens behaved nearly elastically. As the drift loading increased, monolithic Specimen S1
showed a higher equivalent viscous damping ratio than the precast specimens. This was
attributed to the weak interface of monolithic and precast concrete and the slight slip between
the additional bars and concrete in precast specimens at high drift angles. However, when
pinching of the hysteresis loops appeared in the last several drift loading cycles, the equivalent
damping ratio of Specimen S1 was lower than that of Specimen S2. Comparing the two
precast specimens, Specimen S2 was found to have better energy dissipation ability than
Specimen S3, with the relatively higher reinforcement ratio in the beams.

5. Plastic Hinge Length Analysis


One important evaluation index for the ductility of an RC member is plastic hinge length.
A greater plastic hinge length usually indicates better ductility [Paulay and Priestley
1992]. In addition, the plastic hinge zone is the region for strengthening if the member
must be retrofitted and rehabilitated. Therefore, the plastic hinge length is studied here to
estimate the efficiency of the proposed connection. All specimens were tested under cyclic
loading until plastic hinge failure on the beams near the column-to-beam connections.
Plastic hinge failure includes concrete crushing, rebar yielding, and curvature localization
[Jiang et al., 2014; Johnny, 2003; Zhao et al., 2012]. By careful observation of the failure
mode (Fig. 6), it is obvious that the failure area, which is related to the plastic hinge
length, was quite different from S1 to S3. Hence, the effect on the plastic hinge length was
studied here for the sake of the investigation on the mechanical mechanism.

5.1. Plastic Hinge Length of the Specimens


Because the beams were under cyclic loading until plastic hinge failure and the top
reinforcements were different from the bottom reinforcements, the case with the beam
top fibers under compression should be different from the case with the beam top fibers
under tension in plastic hinge length evaluation. Therefore, two situations should be
considered when measuring the experimental plastic hinge length. The plastic deformation
comes from the material yielding of concrete and steel. The region of the steel yielding
zone is very difficult to measure, and the calculated plastic hinge length is more closely
related with the curvature localization zone [Jiang et al., 2014]. It has also been noted that
the pattern and development of primary cracks may affect the plastic hinge length [Zhao,
2012]. Moreover, the concrete cracking region and concrete crushing zone can be mea-
sured by a high-precision Digital Image Correlation (DIC) system analyzing in the
experiments from Fig. 15a, 15b, and 15c. The experimental plastic hinge length was
taken as the average value of the concrete cracking and crushing zone of each situation,
which are listed in Table 4.
18 D. Guan et al.

(a)

(b)

(c)

FIGURE 15 Concrete cracking and crushing zones: (a) S1; (b) S2; and (c) S3 (Unit: mm).

5.2. Discussion of the Plastic Hinge

5.2.1. Effect of Monolithic and Precast Elements. For the loading case with the beam top
fibers under tension, the measured plastic hinge lengths for Specimens S1-S3 are 320.1, 319.0,
and 318.9 mm, respectively. In this case, the tensile reinforcement situations and concrete
TABLE 4 Experimental results of plastic hinge length
Specimen ID S1 S2 S3
Top Bottom Top Bottom Top Bottom
Loading situation tensile tensile tensile tensile tensile tensile
Tensile reinforcement data db or db (mm) 20.0 20.0 20.0 28.7 20.0 26.1
fy or fy (MPa) 448.5 448.5 448.5 447.5 448.5 458.0
As (mm2) 1570.8 942.5 1570.8 1289.6 1570.8 1068.1
Experimental measured results Left beam Concrete cracking 340.6 385.5 569.2 608.4 509.2 584.1
(mm) zone
Concrete crushing 295.0 182.6 125.2 0.0 191.8 0.0
zone
Average 317.8 284.1 347.2 304.2 350.5 292.1
Right beam Concrete cracking 396.3 454.3 323.9 401.1 356.5 506.1
zone
Concrete crushing 248.3 194.2 257.8 127.8 218.1 0.0
zone
Average 322.3 324.3 290.9 264.5 287.3 253.1
Overall average 320.1 304.2 319.0 284.3 318.9 272.6
Development and Seismic Behavior of Precast Beam-column Connections
19
20 D. Guan et al.

strength are very similar. Based on the similar values of the plastic hinge lengths of
approximately 320 mm, it can be concluded that the precast specimens have the same plastic
hinge length of precast beams as the monolithic specimen, and the plastic hinge length of precast
specimens could be calculated using the models for monolithic specimens.

5.2.2. Effect of Tensile Reinforcements. The basic form of Eq. (1) has been widely
adopted and further studied to evaluate the plastic hinge length in recent investigations
[Elmenshawi et al., 2012; Gu et al., 2012; Jiang et al., 2014; Panagiotakos and Fardis,
2001; Paulay and Priestley, 1992].
,

LP ¼ μz þ φfy db ; (1)

where μ and φ are coefficients with different values. The first term indicates the effect of
the moment gradient, and the second term indicates the effect of the strain penetration of
longitudinal reinforcements. In this study, all reinforcements at the beam bottom should
be involved for evaluation when the beam bottom is under tension. If the different
reinforcing bar sizes were arranged (S2 and S3), db is taken as db based on the area
2 P db 2
equivalent: n d4b π ¼ 4 π, where n is the number of tensile reinforcements. If the
n
tensile yielding strengths are different for different reinforcing bars, the average value fy
would be taken. The equivalent yielding strength and diameters of the bars in Specimens
S2 and S3 are also shown in Table 4 to help explain the plastic hinge length. It is
interesting that the effect of strain penetration is related with the yielding strength and
the diameter of the longitudinal reinforcements only. This consideration is for the sake of
the convenience of calculation. However, another control factor of strain penetration was
found in this investigation, which is the total area of the tensile rebar (or reinforcement
ratio). From the situation of beam top fibers under tension to beam bottom fibers under
tension, the reinforcing bar diameter increases, and the total area of the tensile reinfor-
cing bars decreases. Eventually, the calculated results of the beam bottom fibers under
tension would be larger than those of the beam top fibers under tension. However, the
experimental results show an opposite phenomenon. Although the reinforcing bar
diameter is increasing, the experimental plastic hinge length is proportional to the total
area of the tensile reinforcing bars. This factor was proposed by Johnny [2003] for RC
columns. However, for the case of RC beams, the reinforcement ratio factor was also
ignored. Based on the aforementioned novel discovery, a more rational plastic hinge
model for RC beams must be developed in future studies.

6. Conclusions
The newly developed beam-to-column connection combining both longitudinal bar
anchoring and lap splicing used in precast construction has been investigated under
reversal cyclic loading to evaluate seismic performance. Based on the experimental and
analytical investigations, conclusions can be drawn as follows:
1. The precast beam-to-column connections showed strong-column, weak-beam fail-
ure modes, indicating that the strong-column, weak-beam design philosophy was
effective for the proposed system. The precast specimens exhibited fat and stable
hysteretic loops, which represented more satisfactory hysteretic behavior than the
monolithic specimen.
Development and Seismic Behavior of Precast Beam-column Connections 21

2. The strength of precast Specimen S3 was slightly higher than that of the mono-
lithic specimen, so the plane cross-section assumption could be adopted for the
combined system, and the additional bars could be considered in the strength
calculation.
3. Considering the deformability, the precast specimens showed ductile behavior
under reversal cyclic loading. The initial stiffness of the precast specimens was
marginally lower than that of the monolithic specimens. However, after yielding,
the proposed precast connections became slightly stiffer than the monolithic
connection.
4. Although the equivalent viscous damping ratios of precast specimens seemed
slightly lower than that of the monolithic specimen in most loading cycles, the
precast specimens dissipated more overall energy than the monolithic specimen,
showing a satisfactory capacity of energy dissipation.
5. The plastic hinge length of the proposed precast beam-to-column connection could
be calculated by the models for monolithic components. However, most of the
existing models for plastic hinge length ignore the reinforcement ratio, which was
demonstrated to be an important factor in this study.

Funding
The authors gratefully acknowledge financial support from the National “Twelfth Five-Year”
Plan for Science & Technology (No. 2011BAJ10B03) and from a project funded by the
Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD).

References
ACI Committee 374 [2005] “Acceptance criteria for moment frames based on structural testing and
commentary,” ACI 374.1-05, American Concrete Institute, Farmington Hills, MI.
ACI Innovation Task Group 1 and Collaborators (ACI-ITG 1) [2003] “Special hybrid moment
frames composed of discretely jointed precast and post-tensioned concrete members (ACI Tl.2-
03) and commentary (ACI Tl.2R-03),” American Concrete Institute, Farmington Hills, MI.
Alcocer, S. M., Carranza, R., Perez-Navarrete, D., and Martinez, R. [2002] “Seismic tests of beam-
to-column connections in a precast concrete frame,” PCI Journal 47(3), 70–89.
Ali, R. K. and Parastesh, H. [2003] “Cyclic loading of ductile precast concrete beam-column
connection,” ACI Structural Journal 100(3), 291–296.
Applied Technology Council (ATC) [2005] “Improvement of nonlinear static seismic analysis
procedures,” FEMA 440, Federal Emergency Management Agency, Washington, DC.
Cai, J., Zhu, H., Feng, J., Liu, Y., and Huang, L. [2012] “Experimental study on seismic behavior of
middle joints of SCOPE system,” Journal of Central South University: Science and Technology
43(5), 1894–1901. (in Chinese)
Chen, S., Yan, W., and Gao, J. [2012] “Experimental investigation on the seismic performance of
large-Scale interior beam-column joints with composite slab,” Advances in Structural
Engineering 15(7), 1227–1238.
Choi, H. K., Choi, Y. C., and Choi, C. S. [2013] “Development and testing of precast concrete beam-
to-column connections,” Engineering Structures 56, 1820–1835.
Chopra, A. K. [1995] Dynamics of Structures, Prentice Hall, Upper Saddle River, NJ.
Crisafulli, F. and Restrepo, J. [2003] “Ductile steel connections for seismic resistant precast build-
ings,” Journal of Earthquake Engineering 7(4), 541–553.
Elmenshawi, A., Brown, T., and El-Metwally, S. [2012] “Plastic hinge length considering shear
reversal in reinforced concrete elements,” Journal of Earthquake Engineering 16(2), 188–210.
22 D. Guan et al.

El-Sheikh, M., Pessiki, S., Sause, R., and Lu, L. [2000] “Moment rotation behavior of unbonded post-
tensioned precast concrete beam-column connections,” ACI Structural Journal 97(1), 122–131.
El-Sheikh, M., Sause, R., Pessiki, S., and Lu, L. [1999] “Seismic behavior and design of unbonded
post-tensioned precast concrete frames,” PCI Journal 44(3), 54–71.
Ericson, A. [2010] “Emulative detailing in precast concrete systems,” Proceedings of Structures
Congress, Orlando, Florida, pp. 2903–2913.
Ertas, O., Ozden, S., and Ozturan, T. [2006] “Ductile connections in precast concrete moment
resisting frames,” PCI Journal 51(3), 66–76.
fib Bulletin 43 [2008] “Structural connections for precast concrete buildings,” International
Federation for Structural Concrete, Lausanne, Switzerland.
GB50010-2010 [2010] “Code for design of concrete structures,” China Architecture & Building
Press, Beijing, China. (in Chinese)
Gu, D., Wu, Y., Wu, G., and Wu, Z. [2012] “Plastic hinge analysis of FRP confined circular concrete
columns,” Construction and Building Materials 27(1), 223–233.
Guo, Z. X. [2014] “Thinking on generalization and application of new prefabricated concrete
structure,” Construction Technology 43(1), 17–22. (in Chinese)
Ha, S. S., Kim, S. H., Lee, M. S., and Moon, J. H. [2014] “Performance evaluation of semi precast
concrete beam-column connections with U-shaped strands,” Advances in Structural Engineering
17(11), 1585–1600.
Hosoya, H., Kimura, T., Knankubo, T., and Yasojima, A. [2012] “Experimental study on structural
performance of precast RC beam-column joints,” AIJ Journal of Technology and Design 18(39),
917–922. (in Japanese)
Im, H. J., Park, H. G., and Eom, T. S. [2013] “Cyclic loading test for reinforced-concrete-emulated
beam-column connection of precast concrete moment frame,” ACI Structural Journal 110(1),
115–125.
Jiang, C., Wu, Y., and Wu, G. [2014] “Plastic hinge length of FRP-confined square RC columns,”
Journal of Composites for Construction 18(4), 04014003.
Johnny, H. C. M. [2003] “Inelastic design of reinforced concrete beams and limited ductile high-
strength concrete columns,” Ph.D. thesis, The University of Hong Kong, Hong Kong.
Korkmaz, H. H. and Tankut, T. [2005] “Performance of a precast concrete beam-to-beam connection
subject to reversed cyclic loading,” Engineering Structures 27(9), 1392–1407.
Le-Trung, K., Lee, K., Shin, M., and Lee, J. [2013] “Seismic performance evaluation of RC beam-
column connections in special and intermediate moment frames,” Journal of Earthquake
Engineering 17(2), 187–208.
Li, B., Kulkarni, S. A., and Leong, C. L. [2009] “Seismic performance of precast hybrid-steel
concrete connections,” Journal of Earthquake Engineering 13(5), 667–689.
Lu, X. Z., Teng, J. G., Ye, L. P., and Jiang, J. J. [2005] “Bond–slip models for FRP sheets/plates
bonded to concrete,” Engineering Structures 27(6), 920–937.
Maya, L., Zanuy, C., Albajar, L., Lopez, C., and Portabella, J. [2013] “Experimental assessment of
connections for precast concrete frames using ultra high performance fibre reinforced concrete,”
Construction and Building Materials 48, 173–186.
Pampanin, S., Priestley, M. N., and Sritharan, S. [2001] “Analytical modelling of the seismic
behaviour of precast concrete frames designed with ductile connections,” Journal of
Earthquake Engineering 5(3), 329–367.
Panagiotakos, T. B. and Fardis, M. N. [2001] “Deformations of reinforced concrete members at
yielding and ultimate,” ACI Structural Journal 98(2), 135–148.
Parastesh, H., Hajirasouliha, I., and Ramezani, R. [2014] “A new ductile moment-resisting connec-
tion for precast concrete frames in seismic regions: An experimental investigation,” Engineering
Structures 70, 144–157.
Park, R. [1995] “A perspective on the seismic design of precast concrete structures in New Zealand,”
PCI Journal 40(3), 40–60.
Park, R. and Bull, D. K. [1986] “Seismic resistance of frames incorporating precast prestressed
concrete beam shells,” Journal Prestressed Concrete Institute 31(4), 54–93.
Development and Seismic Behavior of Precast Beam-column Connections 23

Paulay, T. and Priestley, M. [1992] Seismic Design of Reinforced Concrete and Masonry Buildings,
John Wiley & Sons, New York.
Restrepo, J. I., Park, R., and Buchanan, A. H. [1995] “Tests on connections of earthquake resisting
precast reinforced-concrete perimeter frames of buildings,” PCI Journal 40(4), 44–61.
Shariatmadar, H. and Zamani, E. B. [2014] “An investigation of seismic response of precast concrete
beam to column connections: Experimental study,” Asian Journal of Civil Engineering-Building
And Housing 15(1), 41–59.
Sucuoglu, H. [1995] “Effect of connection rigidity on seismic response of precast concrete frames,”
PCI Journal 40(1), 94–103.
Xue, W. C. and Yang, X. L. [2010] “Seismic tests of precast concrete, moment-resisting frames and
connections,” PCI Journal 55(3), 102–121.
Yee, A. A. [2001] “Social and environmental benefits of precast concrete technology,” PCI Journal
46(3), 14–19.
Zhao, X. [2012] “Investigation of plastic hinges in reinforced concrete (RC) structures by finite
element method and experimental study,” Ph.D. thesis, The University of Hong Kong, Hong
Kong.
Zhao, X., Wu, Y., and Leung, A. [2012] “Analyses of plastic hinge regions in reinforced concrete
beams under monotonic loading,” Engineering Structures 34, 466–482.

View publication stats

Potrebbero piacerti anche