You are on page 1of 24

EARTHQUAKE ENGINEERING AND STRUCTURAL DYNAMICS

Earthquake Engng Struct. Dyn. 2009; 38:331–354


Published online 6 November 2008 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/eqe.857

Quasi-static and pseudo-dynamic testing of unbonded


post-tensioned rocking bridge piers with external
replaceable dissipaters

Dion Marriott1, ∗, † , Stefano Pampanin1 and Alessandro Palermo2


1 Department of Civil and Natural Resources Engineering, University of Canterbury, Christchurch, New Zealand
2 Department of Structural Engineering, Politecnico di Milano, Italy

SUMMARY
It has been well documented that following a major earthquake a substantial percentage of economic
loss results from downtime of essential lifelines in and out of major urban centres. This has thus led
to an improvement of both performance-based seismic design philosophies and to the development of
cost-effective seismic structural systems capable of guaranteeing a high level of protection, low structural
damage and reduced downtime after a design-level seismic event. An example of such technology is the
development of unbonded post-tensioned techniques in combination with rocking–dissipating connections.
In this contribution, further advances in the development of high-performance seismic-resistant bridge
piers are achieved through the experimental validation of unbonded post-tensioned bridge piers with
external, fully replaceable, mild steel hysteretic dissipaters.
The experimental response of three 1 : 3 scale unbonded, post-tensioned cantilever bridge piers, subjected
to quasi-static and pseudo-dynamic loading protocols, are presented and compared with an equivalently
reinforced monolithic benchmark. Minimal physical damage is observed for the post-tensioned systems,
which exhibit very stable energy dissipation and re-centring properties. Furthermore, the external dissipaters
can be easily replaced if severely damaged under a major (higher than expected) earthquake event. Thus,
negligible residual deformations, limited repair costs and downtime can be achieved for critical lifeline
components. Satisfactory analytical–experimental comparisons are also presented as a further confirmation
of the reliability of the design procedure and of the modelling techniques. Copyright q 2008 John Wiley
& Sons, Ltd.

Received 14 April 2008; Revised 28 August 2008; Accepted 29 August 2008

KEY WORDS: bridge piers; unbonded post-tensioning; cyclic tests; pseudo-dynamic; external replaceable
dissipaters; energy dissipation

∗ Correspondence to: Dion Marriott, Department of Civil and Natural Resources Engineering, University of Canterbury,
Christchurch, New Zealand.

E-mail: djm178@student.canterbury.ac.nz

Copyright q 2008 John Wiley & Sons, Ltd.


332 D. MARRIOTT, S. PAMPANIN AND A. PALERMO

INTRODUCTION

In recent literature the performance of unbonded post-tensioned structures undergoing controlled


rocking at the base has highlighted an improved structural performance when compared with
their equivalently reinforced monolithic counterparts, both for buildings [1–3] and bridge systems
[4–6]. For controlled rocking structures, inelastic deformation is lumped to a number of discrete
rocking connections reducing damage to the adjacent structural elements, while unbonded post-
tensioned tendons return the structure to its original position, minimizing residual (or permanent)
deformations.
A number of experimental and analytical studies have investigated the use of either unbonded or
partially bonded prestressed reinforcement for use in precast bridge pier construction [5, 7–11].
Further studies have also investigated the pseudo-dynamic response of single-degree-of-freedom
(SDOF) bridge piers [12] or multi-degree-of-freedom (MDOF) bridge systems using non-linear
substructure modelling algorithms [13].
The focus of this paper is on the development and testing of a cost-effective system capable of
achieving a high level of performance under a design-level earthquake event or greater (500 year
return period or greater). To achieve this objective, replaceable energy dissipation devices, based
on axial yielding of mild steel elements, are combined with unbonded post-tensioned techniques.
Direct costs associated with repair and indirect costs associated with downtime are drastically
minimized after a major event due to the immediate replacement of the devices and lack of
structural damage to the precast elements.
The results from a series of quasi-static and pseudo-dynamic tests on four 13 scaled cantilever
bridge pier specimens, namely one unbonded post-tensioned only specimen, two hybrid specimens
(having a combination of unbonded post-tensioned tendons and replaceable dissipaters) and one
monolithic benchmark pier, are presented.
A description of the analytical macro-model used to represent the hybrid pier is also given.
A comparison with the experimental results confirmed the reliability of the adopted modelling
technique based on a refined macro-model.

DEVELOPMENT OF HYBRID SYSTEMS: USE OF EXTERNAL


REPLACEABLE DISSIPATERS

Traditional hybrid systems, consisting of unbonded post-tensioned tendons, combined with mild
steel reinforcement, primarily comprise of internally grouted reinforcement either for beam–column
connections [1, 14], structural walls [15], or for bridge pier systems [6, 10]. More recently, the
focus has been on the development of hybrid systems with external dissipaters in order to reinstate
the structural integrity of the system following a major event with minimal downtime. Furthermore,
the use of external dissipaters provides greater efficiency as the system does not suffer any stiffness
degradation as a result of possible bond deterioration, typical of internally grouted reinforcement
[10]. From a design point of view, greater flexibility exists as the designer can refine the design
by adopting a specific number of non-standard bar diameters manufactured from readily available
standard mild steel bar. The prototype bridge pier in Figure 1 illustrates the two hybrid concepts
for a SDOF bridge pier system (internal versus external dissipation), where the rocking behaviour
is located at the foundation–pier interface. In this example, the external dissipaters could in fact

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
UNBONDED POST-TENSIONED ROCKING BRIDGE PIERS 333

Figure 1. Prototype bridge pier and dissipation options for rocking systems.

comprise of mild steel reinforcement, fluid viscous dampers, friction dampers, Shape Memory
Alloy devices or a combination of the above.
As a part of an extensive research programme on the development of high-performance seismic-
resisting bridge pier systems dealing specifically with unbonded post-tensioned techniques at the
University of Canterbury, the experimental results of four bridge pier specimens are presented
herein. Preliminary results on similar bridge piers with internally grouted mild steel reinforcement
(non-replaceable) subjected to quasi-static cyclic testing have been presented in [10, 16].
In this contribution the focus is on two hybrid specimens, adopting replaceable external mild steel
dissipaters, tested under uni-directional quasi-static and pseudo-dynamic loading. The experimental
results of a single monolithic benchmark and an unbonded post-tensioned only system are presented
for comparison purposes alone.

DESIGN OF THE PROTOTYPE BRIDGE PIER

The dimensions of the prototype bridge pier are presented in Figure 1, having a total participating
weight of 1800 kN with a cantilever height of 4.8 m and square cross-section dimension of 1.05 m.
The test specimens were a 1 : 3 scale representation and therefore required a total weight of 200 kN,
a cantilever length of 1.6 m and square dimension of 0.35 m (constant mass density used during
the scaling process).
A monolithic bridge (conventional ductile piers), defining the prototype system, was designed
according to a direct displacement-based design (DDBD) methodology [17]. The DDBD uses
an equivalent elastic SDOF substitute structure with secant stiffness to the target displacement.
The target displacement was, in this case, the main performance objective for design targeting a
lateral drift of 2%. The design spectrum was reduced based on a displacement ductility of  = 3.8
(converted to equivalent viscous damping), requiring a strength-to-mass ratio of 0.3 g. This equated
to a base shear of 60 kN at a lateral displacement of 32 mm for the scaled test specimen.
A benchmark monolithic pier, denoted as MON1, was designed to achieve the performance
objective stated above. Two hybrid piers were designed to have similar objectives, denoted as

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
334 D. MARRIOTT, S. PAMPANIN AND A. PALERMO

HBD3 and HBD4. The hybrid pier HBD4 was designed to achieve the same performance objective
as the monolithic; that is, a lateral strength of 60 kN at 32 mm. In contrast, the hybrid pier, HBD3,
was designed to be under strength by approximately 15%. Alternatively, the hybrid pier, HDB4,
could have been independently designed following a DDBD for hybrid systems [6]. However,
when compared with the monolithic pier, the hybrid pier would require greater strength (at the
same target displacement). For comparison sake, it was decided to design both the benchmark
(MON1) and the hybrid (HBD4) to achieve the same strength at a lateral displacement of 32 mm.
A second performance objective was included within the design of the hybrid piers. This objective
required that the tendons remain elastic and rupture of the non-prestressed reinforcement was to
be avoided under a maximum considered event (MCE). The MCE lateral drift demand was taken
as 3.5% (spectral ordinates amplified by R = 1.5). Further details outlining the design objectives,
performance objectives, and specimen details can be found elsewhere [16].

DETAILS OF THE BRIDGE PIER TEST SPECIMENS

The geometry and reinforcement details for the two hybrid pier specimens (HBD3 and HBD4),
the unbonded post-tensioned only pier (PT) and the equivalently reinforced monolithic counterpart
(MON1) are summarized in Table I. The testing regime for each pier specimen is also presented
in Table I. Uni-directional quasi-static testing was carried out on each of the four test specimens
(MON1, PT, HBD3, HBD4). In addition to this, pseudo-dynamic testing for two recorded strong
ground motions was performed on the hybrid specimen HBD3 as well as on the benchmark MON1.
The details of the testing programme are discussed in the following section.

Bridge pier construction details—general


The monolithic benchmark bridge pier (MON1) required 16-D10 (grade 300 MPa, 10 mm reinforce-
ment) symmetrically located about the perimeter of the section (Figure 2: left). The construction
details specific to the two hybrid pier sections (HBD3 and HBD4) are shown in Figure 3, while
the dissipater connection detail is illustrated in Figure 4.
Both hybrid systems had the same initial post-tensioning and prestressing reinforcement content;
however, HBD3 had less mild steel dissipation than HBD4. The reduction in the lateral capacity
of HBD3, as discussed above, was chosen to highlight the greater displacement capacity inherent
within unbonded post-tensioned rocking systems; this would become evident with respect to the
pseudo-dynamic testing where larger displacements can be anticipated, but also well accommodated
(Table II).

Construction details of the hybrid pier HBD3


Four unbonded post-tensioned tendons were located in the centre of HBD3 (and HBD4), where
each tendon was initially post-tensioned to 75 kN, resulting in a total initial prestressing force
of 300 kN (Figure 3: top right and Figure 4: right). Dissipation was supplied via four external
replaceable dissipaters, whose conceptual development and behaviour are discussed later. The
four dissipaters were located on two faces; two dissipaters per face in the direction of loading
(as opposed to be being located at each end of the pier section). Each mild steel dissipater was
fabricated from a standard mild steel bar (300 MPa) and turned in a lathe to have a fuse diameter of

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
Copyright q
Table I. Test specimen details.
Specimen Monolithic MON1 Post-tensioned PT Hybrid HBD3 Hybrid HBD4

2008 John Wiley & Sons, Ltd.


Geometry 350×350 mm square 350×350 mm square 350×350 mm square 350×350 mm square

Axial load/initial 200 kN—constant axial load 300 kN—4 tendons at 75 kN 300 kN—4 tendons at 75 kN 300 kN—4 tendons at 75 kN
post-tensioning control each each each

Mild steel Cast-in-place construction: — 4–10 mm diameter mild 8–8 mm diameter mild steel
reinforcement/ 16-D10 grade 300 MPa steel (Grade 300 MPa) (Grade 300 MPa) machined
dissipation devices reinforcing. machined dissipaters. 75 mm dissipaters. 115 mm machined
machined fuse length. fuse length.

Testing regime Quasi-static, Quasi-static Quasi-static, Quasi-static


pseudo-dynamic pseudo-dynamic
UNBONDED POST-TENSIONED ROCKING BRIDGE PIERS

DOI: 10.1002/eqe
Earthquake Engng Struct. Dyn. 2009; 38:331–354
335
336 D. MARRIOTT, S. PAMPANIN AND A. PALERMO

Figure 2. Monolithic pier details; left: geometry and reinforcement, right: construction.

Figure 3. Hybrid pier specimen details; left: pier element (without dissipation); centre: pier Section A-A;
right top: HBD3; right bottom: HBD4.

10 mm over a gauge length equal to 75 mm. This length was chosen to prevent premature rupture
of the mild steel.

Construction details of the hybrid pier HBD4


The hybrid specimen HBD4 (Figure 3: bottom right and Figure 4: right) was identical to HBD3
albeit for a greater number of dissipater elements (eight external replaceable dissipaters for HBD4,
versus four dissipaters for HBD3). Details relating to the dissipater arrangement for HBD3 and
HBD4 are presented in Figures 3 and 4. For HBD4 to achieve the same capacity as the monolithic

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
UNBONDED POST-TENSIONED ROCKING BRIDGE PIERS 337

Figure 4. Dissipater attachment details; left: as built specimen; right: schematic comparison.

Table II. Earthquake records selected for pseudo-dynamic testing.


Ref Earthquake event Year Station Soil type (NZS1170) Scaled PGA (g)
cm1 Cape Mendocino 1992 Fortuna Blvd C 0.340
lp5 Loma Prieta 1989 Hollister Diff Array B 0.300

pier MON1, the mild steel dissipaters were turned in the lathe to have a fuse diameter of 8 mm over
a fused length equal to 115 mm. A longer fuse length was required for HBD4, as the dissipaters
were located at each end of the section and were thus subjected to larger displacement demands.
Two dissipaters were located on each face of HBD4 resulting in a symmetrical distribution for the
dissipaters. Both HBD3 and HBD4 comprise of the same precast pier element, only differing in
the amount (and location) of external dissipaters that were installed prior to testing.
The dissipaters were attached to the precast pier element via an externally mounted steel bracket
(see Figures 3 and 4). Four M25 high-strength threaded rods pass through the pier section in each
direction, fixing one steel bracket to each of the four pier faces. To further prevent sliding of the
steel plate on the concrete, the surface of the steel and concrete was roughened and a film of heavy
duty epoxy was applied to both surfaces. This allowed the dissipater forces to be transferred to
the pier element via friction and mechanical interlock. The foundation end of dissipater is bolted
to a steel block attached to the top of the foundation. The steel block is bolted to the foundation
with two M12, 880 MPa threaded rods, which are epoxy injected into the concrete foundation to
an embedment depth of 250 mm (refer Figure 3).
On the top surface of the precast concrete foundation (the rocking interface) a 15 mm steel plate
was cast in situ. A 20 mm steel plate was also cast to the underside of the precast pier and welded
to the longitudinal reinforcement. This would result in a steel–steel rocking interface and prevent
damage to the cover concrete.
Finally, shear transfer between the pier and foundation was sufficiently accounted for through
the use of a metallic spherical ball (foundation) and socket (pier underside) located at the centre
of the pier section (Figure 6, inset).

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
338 D. MARRIOTT, S. PAMPANIN AND A. PALERMO

QUASI-STATIC AND PSEUDO-DYNAMIC TESTING

The experimental set-up was identical for both the hybrid and monolithic pier specimens (Figure 5).
Specific instrumentation schemes were adopted depending on whether rotation or curvature was to
be measured for each test. During testing of the monolithic benchmark pier (MON1) the 200 kN
axial load, representing the gravity deck loading, was kept constant. For the hybrid specimens,
an axial load of 300 kN, representing the gravity deck loading and initial post-tensioning, would
increase as a result of tendon elongation due to the opening of the gap at the foundation level.
The quasi-static loading regime consisted of three cycles at each drift level, followed by a single
intermediate cycle (Figure 6). The procedure defining the loading protocol is adopted from the
ACI recommendations ‘Acceptance criteria for moment frames based on structural testing’ [18].
This standard defines the minimum experimental evidence required in order to construct structural
systems outside the requirements of ACI 318-99.
The document provides a number of prescriptive pass/fail criteria. On completion of the 3rd
and final cycle at an inter-storey drift level not less than 3.5% the system must satisfy the

Figure 5. Experimental test set-up; left: monolithic specimen; right: hybrid specimen.

Figure 6. Photo of experimental test set-up for PT test (left), quasi-static, displacement loading protocol
(right), ball and socket shear key mechanism (bottom left inset).

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
UNBONDED POST-TENSIONED ROCKING BRIDGE PIERS 339

following:
1. The peak force in any given loading direction must exceed 75% of the maximum force for
the same loading direction.
2. The relative energy dissipation (ratio of energy dissipated to an equivalent elasto-plastic
system) shall exceed 1 : 8 (0.1275).
3. The secant stiffness from −3.5 to 3.5% of lateral drift shall exceed 0.05 times the initial stiff-
ness of the system (determined from the elastic cycles at the beginning of the testing protocol).
The pseudo-dynamic testing was performed under two recorded strong ground motions. A
Newmark integration algorithm was used to solve the SDOF equation of motion. The damping
model implemented in the integration algorithm consisted of initial stiffness proportional damping
and was set equal to 5% of critical damping.
It is worth noting that recent studies [19] have shown that an initial stiffness damping model can
lead to a non-conservative design as it overestimates the intrinsic damping forces associated with
the elastic response. A damping model proportional to the tangent stiffness has been suggested as
a more correct alternative. Considering that the main aims of the pseudo-dynamic testing were:
(a) to confirm the performance of a hybrid system having a reduced level of damage and residual
deformations under a design-level event and (b) to further validate the analytical model, the adopted
damping model (proportional to initial stiffness) can be considered more than adequate for the
scope of this investigation.
The earthquake records were selected according to seismological characteristics (soil type,
source-to-site distance and intensity) and scaled according to the New Zealand seismic loading
code [20] for a PGA of 0.58g located on shallow soil (soil category C) as shown in Figure 7.
For this investigation the earthquake records were scaled over a period range Teff to 0.707Teff
defined by a ductility range of 0.5target to target . Within a DDBD, the effective period is defined
by the secant stiffness to the target displacement (defining the target ductility, target ). As a result

Figure 7. Scaled acceleration time history records for pseudo-dynamic testing and their
corresponding elastic response spectra.

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
340 D. MARRIOTT, S. PAMPANIN AND A. PALERMO

of the low design ductility, this would actually result in a relatively small window over which to
scale the records to.

DEVELOPMENT AND TESTING OF EXTERNAL REPLACEABLE DISSIPATERS

The key feature of the hybrid configuration (unbonded post-tensioned tendons plus dissipaters)
presented in this contribution is given by the implementation of cost-efficient replaceable dissi-
paters. The dissipaters are axial, tension–compression yielding mild steel bar elements, restrained
against buckling and mounted external to the pier section. The connection detail was appropriately
detailed to allow the dissipaters to be removed after testing (or after an earthquake, as per the
prototype structure).
The dissipaters were manufactured in the Civil Engineering Laboratory at the University of
Canterbury from 20 mm diameter mild steel bar and provided with a reduced diameter over a
specified fuse length. The fabrication process is illustrated in Figure 8. The bar is first placed in
a lathe and ‘turned’ to the desired diameter (the two dissipater sets were turned to 10 and 8 mm
for their specific applications presented herein). Next, the steel dissipater is strain gauged (for
instrumentation purposes only in the test specimen). A 34 mm (outside diameter) steel tube, with a
wall thickness of 2 mm, is located over the machined area of the steel bar and temporarily fixed in
place. Epoxy is then injected into a hole provided at the bottom of the steel tube to ensure all the
air is expelled out of the opening at the top. This epoxy steel-tube system prevents buckling of the
machined length and allows the dissipater to yield efficiently in compression. The manufacturing
process is relatively simple and cost effective, the most time consuming and expensive phase being
the strain gauging: this is clearly not required in practice.
As the dissipaters are located outside of the pier section, consideration should be given to
environmental weathering; the use of protective coatings similar to that currently employed in
structural steel applications would seem appropriate.
Prior to their installation within the bridge pier, the individual dissipaters were subjected to
cyclic, axial, tension–compression tests to characterize their cyclic energy dissipation and stability;
particularly, with respect to the capability of the anti-buckling system. A typical dissipater test

Figure 8. Manufacturing process of the mild steel dissipaters; left: turning the mild steel bar in the lathe,
centre left: strain gauging (only required for testing), centre right: installation of the anti-buckling tube.
Right: epoxy injection of the anti-buckling tube.

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
UNBONDED POST-TENSIONED ROCKING BRIDGE PIERS 341

Figure 9. Testing of the mild steel dissipaters; left: force–displacement response


(rupture highlighted), centre: test set-up, top right: failure mechanism, bottom right:
displacement testing protocol (failure highlighted).

set-up is shown in Figure 9 (centre). The axial elongation is measured over a gauge length
equivalent to that being provided within the bridge pier specimen. The test was carried out with
a displacement protocol compatible with what was expected during actual testing of the test pier
specimens (Figure 9, bottom right). The cyclic response of a typical dissipater is shown in Figure 9
(left) in terms of axial load versus displacement. In addition, overlain in the same figure are three
monotonic tension tests carried out from the same mild steel bar.
The dissipater was capable of providing a large amount of dependable energy dissipation.
Furthermore, no pinching was observed and the response was very stable, indicating that the anti-
buckling system was working very well. When analysing the failure mechanism after testing, the
effect of the epoxy-tube system was clear; very minor buckling (of a high order) was observed
(Figure 9, top right), resulting in very high, stable energy dissipation as the steel could yield
effectively in tension and compression. It can be seen that on unloading from large displacements
and on approaching the origin, the stiffness actually begins to increase possibly due to the partially
buckled regions bearing on the surface of the epoxy wall.
It is immediately clear that due to a combined fatigue-buckling phenomenon, the average ultimate
strain is significantly reduced when subjecting the steel element to cyclic load reversals (when
compared with monotonic tension testing). Rupture of the dissipater above occurred during the third
cycle approaching an elongation of 10 mm; this is equivalent to a strain of approximately 133×
10−3 mm/mm (whereas, for the monotonic tension tests strains exceeded 200×10−3 mm/mm).
Traditional damage indices recognize that low cycle fatigue reduces the allowable deformation
capacity of an element due to cumulative inelastic energy dissipation, increasing damage to the
structural element [21–24]. It would be possible to calibrate the steel dissipater tests to a number of
local damage indices; however, there is not a 1 : 1 correlation between the local dissipater damage
and global structural damage, i.e. failure of the dissipater does not result in structural failure for
hybrid systems where significant redundancy exists due to the presence of unbonded post-tensioned
tendons. In fact, a weighted damage index based on local damage indices of the post-tensioned
tendons, and mild steel dissipaters would be more appropriate.

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
342 D. MARRIOTT, S. PAMPANIN AND A. PALERMO

Although not presented here for clarity, a simple analytical model based on available steel
hysteresis rules [25] can accurately capture the entire dissipater response observed during testing.
More details of the modelling of such dissipater configurations can be found in [16].

QUASI-STATIC EXPERIMENTAL RESULTS

Experimental response of the monolithic benchmark pier (MON1)


The experimental response of the monolithic benchmark test, designed according to NZS3101
(2006), is presented in Figure 10. Typical of well-designed and detailed specimens, the response
is very stable with large inelastic cycles delivering a large amount of energy dissipation and
consequently large static residual deformations. Significant pinching and stiffness degradation
results from bond deterioration of the cast in situ mild steel reinforcement and only a few large
cracks developed as a result of casting the pier element in two stages (foundation, then the pier, as
per typical construction practice). No significant loss in strength was observed for lateral drifts up
to 4.5%. As mentioned, damage was confined to a few wide cracks. The wider cracks were located
predominantly at the pier–foundation interface and to a lesser extent up a height approximately
equal to the depth of the pier (350 mm). Owing to the use of high-strength concrete ( f c = 65.9 MPa),
spalling was not observed until drifts exceeded 4.5%. The intersection of the horizontal and
vertical-dashed grey lines in Figure 10 indicates the design objective (a lateral load of 60 kN at a
lateral displacement of 32 mm); these co-ordinates compare well with the loading envelope.

Experimental response of the post-tensioned only pier (PT)


The unbonded post-tensioned only pier (PT) can be characterized by a stable non-linear elastic
force–displacement response in Figure 11. The response is extremely stable as the anchorages
of the post-tensioned tendons were preloaded prior to testing to minimize the prestressing losses
associated with anchorage take-up. No damage was evident after testing the specimen to a lateral
drift of 3.5%. This is further shown in that no hysteretic energy dissipation is captured during

Figure 10. Monolithic benchmark; left: experimental force displacement, right: photos
showing damage at 3.5% lateral drift.

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
UNBONDED POST-TENSIONED ROCKING BRIDGE PIERS 343

Figure 11. Unbonded post-tensioned pier, left: experimental force displacement,


right: damage at 3.5% lateral drift.

Figure 12. Left: HBD3 experimental force–displacement response, right: experimental


photos (uplift of dissipaters highlighted).

testing. The pier suffered only minor flexural cracks, which upon unloading, were no longer visible
as a result of the prestressing clamping force.

Experimental response of the hybrid piers (HBD3 and HBD4)


The experimental force–displacement response of both hybrid specimens, HBD3 and HBD4, is
illustrated in Figures 12 and 13. The hybrid pier, HBD3, indicated a very stable hysteretic response
with negligible stiffness degradation and static residual deformations. Damage was limited to
flexural cracking along the pier element close to the rocking interface. Upon unloading these
cracks closed and became unnoticeable. The imposed displacement demands acting on the steel

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
344 D. MARRIOTT, S. PAMPANIN AND A. PALERMO

Figure 13. Left: HBD4 experimental force–displacement response, right: experimental


photos (uplift of dissipaters highlighted).

dissipaters at a lateral drift of 3.5% are in the order of 6.5 mm and indicated in Figure 12 (right).
The inelastic deformation is thus concentrated to these sacrificial and replaceable dissipaters.
The externally mounted connections provided essentially no slip, maximizing the efficiency of
the dissipaters. Furthermore, rupture of the dissipaters did not occur and the tendons remain elastic
up to 3.5% of lateral drift, satisfying the MCE collapse prevention objective. All three of the ACI
acceptance criteria were met, with a relative energy dissipation ratio equal to 0.16 (>0.1275) for
the final cycle and a secant-to-initial stiffness ratio equal to 0.22 at 3.5% lateral drift.
The dashed grey lines in Figure 12, representing the 0.67 MCE design objective, indicate that
the strength of HBD3 (at 32 mm of lateral displacement) did not meet the required design objective.
In fact, this was expected, but investigated to highlight the performance of post-tensioned systems
through subsequent pseudo-dynamic testing in the following section.
The experimental response of the hybrid pier HBD4, in Figure 13, also provided a stable
response. Premature rupture of one dissipater occurs in each direction of loading during the final
load–displacement cycle approaching 3.5% lateral drift. It is worth noting that this rupture was
mainly due to a slightly inadequate dissipater length of 115 mm, due to a physical length constraint
within the connection. While monotonic tensile testing of the mild steel bars indicated available
mild steel strain capacities in the order of 200×10−3 mm/mm, cyclic testing suggested that the
available strain capacity can be reduced by as much as 35–45% due to low cycle fatigue (refer
Figure 9). Multiple cyclic dissipater tests revealed that the strain at rupture typically occurred
in the region of 100×10−3 –140×10−3 mm/mm. The displacement demands on the dissipaters,
located at the edge of the HBD4 pier section, approached 14 mm over a gauge length of 115 mm
at a lateral drift of 3.5%. This corresponds to an average strain of 120×10−3 mm/mm so it is of
no surprise that rupture to one or more of the mild steel dissipaters occurred during testing.
The specified collapse prevention objective at the MCE limit state was, therefore, not completely
satisfied for HBD4 as two out of a total of eight dissipaters ruptured; however, given the high
redundancy of the system, satisfactory performance was still achieved. It should be recalled that at

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
UNBONDED POST-TENSIONED ROCKING BRIDGE PIERS 345

the MCE level the intent is to avoid structural collapse. The vertical load-bearing capacity can be
well maintained after a few dissipaters fail as the post-tensioned tendons are still fully contributing
to the lateral capacity (in fact, the tendons will, in general, contribute to at least 50% of the lateral
capacity).
The higher percentage of energy dissipation is clearly evident for HBD4 when compared with
HBD3, resulting in minor static residual deformations. As reported for HBD3, the damage to
HBD4 was limited to flexural cracking of the pier and inelastic yielding of the steel dissipaters.
For HBD4 all three ACI acceptance criteria were met, with a 10% loss in strength occurring as
a result of rupture to one dissipater in both directions, relative energy dissipation ratio equal to
0.27 (>0.1275) and a secant-to-initial stiffness ratio of 0.14 at 3.5% lateral drift. Furthermore,
the experimental loading envelope of HBD4 agrees very well with the design objective of 60 kN
lateral load at 32 mm of lateral displacement (dashed grey lines in Figure 13).

Consideration of strain limits adopted in the design standards


It is worth noting that the current version of Appendix B of the New Zealand concrete standard,
‘Special provisions for the seismic design of ductile jointed precast concrete structural systems’
[26], suggests that the strain in the non-prestressed mild steel element to be limited to values lower
than 90% of the ultimate strain at the design-level drift (i.e. s 0.9u ) without defining how to
evaluate the ultimate strain. This would apparently suggest that, as this strain limit applies to the
‘design-level drift’ (e.g. 23 MCE, 500 year return period or 2–2.5% lateral drift), there is almost
no reserve deformation capacity for the maximum considered event (i.e. MCE, 2500 year return
period or 3.5–4% lateral drift). In reality, a reduced or typically accepted strain limit of 5–6%
(0.05–0.06 mm/mm) should be adopted as an ‘allowable design strain capacity’.
The ACI concrete standard of [27] gives similar guidance for the ultimate steel strain. In this
case, the maximum allowable strain of 0.9u relates to a story drift equal to or greater than 3.5%,
which is understood to be a maximum considered displacement demand (further confirmed in
other publications [18]). The ultimate strain, however, is still not explicitly defined, requiring
information from measured (or known) steel stress–strain relationships, which may or may not
have any allowance for cyclic loading.
It could be argued that the New Zealand guidelines are more relaxed, permitting a performance
objective where rupture of the non-prestressed reinforcement is tolerated at the maximum credible
event. After all, in most cases the MCE will correspond to a collapse prevention limit state, not a
damage control limit state; rupture of some of the mild steel would not comprise the overall lateral
stability of the system having at least a 50% residual strength being provided by the prestressed
reinforcement. Furthermore, when considering the replaceable nature of external dissipaters, the
New Zealand guidelines can be further justified.

PSEUDO-DYNAMIC EXPERIMENTAL RESULTS

Monolithic pier specimen (MON1)


The displacement performance objectives were, in general, met for the monolithic specimen where
lateral displacements of 28.4 mm (1.8% lateral drift) and 27.7 mm (1.7% lateral drift), respectively,
were recorded (Figure 14). The damage observed to MON1 was confined to a few larger cracks
at the base of the pier as opposed to a distributed plastic hinge region (Figure 15). As discussed

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
346 D. MARRIOTT, S. PAMPANIN AND A. PALERMO

Figure 14. Pseudo-dynamic experimental results for MON1; top: cm1 record, bottom: lp5 record.

earlier, this was a result of the cold joint during construction as a result of two separate concrete
pours. For the same reason, residual deformations registered after the tests are not significant.
The ability of the monolithic specimen to provide continued efficient dissipation has been
significantly reduced following each earthquake; the hybrid specimen on the other hand will
continue to have stable, efficient dissipation until the dissipaters rupture (regardless of whether the
dissipaters are replaced or not). The softening and reduction in hysteretic energy dissipation, due
to bond deterioration, is clearly evident during the second pseudo-dynamic test (in this case being
used to represent a sort of aftershock).
Based on ongoing numerical studies [28], it could be anticipated that a near-field record with a
single large velocity pulse in one direction would have drastically increased the residual deforma-
tions of the monolithic specimen; whereas, for the same ground motions the residual deformations
of the hybrid piers would be negligible.

Hybrid pier specimen (HBD3)


Pseudo-dynamic testing of the hybrid system revealed greater stability and performance when
compared to the monolithic specimen. The maximum recorded displacements were 34 mm (2.1%
lateral drift) and 30 mm (1.9% lateral drift) under both ground motions, respectively (Figure 16).

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
UNBONDED POST-TENSIONED ROCKING BRIDGE PIERS 347

Figure 15. Damage to MON1 after pseudo-dynamic testing.

Considering that the lateral capacity of the hybrid specimen (HBD3) was 15% lower than the
monolithic pier, the maximum displacements of the hybrid pier were actually very comparable to
the monolithic pier. On average, a lateral drift of 2.0% was recorded for HBD3 versus 1.8% for
MON1. The damage sustained to HBD3 was minimal, consisting of minor flexural cracking and
inelastic yielding of the external dissipaters alone (without rupture).
A concluding remark should be noted regarding the pseudo-dynamic testing. As the damping
model uses an initial stiffness proportional damping model, in addition to a relatively high level
of critical damping (5%), the displacements may in fact be artificially (and non-conservatively)
lower than the targeted displacement adopted during the displacement-based design procedure. In
fact, previous numerical studies published in [29] and referring to a monolithic specimen have
indicated that a 15–20% increase in the maximum displacement response can be expected if a
tangent stiffness proportional damping model is implemented when compared with a damping
model proportional to the initial stiffness (considering 5% critical damping for each damping
model).

NUMERICAL MODEL

A wide range of modelling techniques exist in practice, and in the research community, capable of
modelling the response of controlled rocking systems ranging from 3-dimensional finite element
models (3D-FEM), fibre models or simple, yet sufficiently accurate, macro-models. An overview
of modelling techniques can be found in [30, 31]. With regards to a macro-model, despite the
reduced computational effect, a good physical understanding of the problem is required in order
to correctly implement the model, and the associated parameters, and to be confident with the
output. The efficiency of two macro-models, both based on a lumped plasticity concept, is the
focus within this research project.

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
348 D. MARRIOTT, S. PAMPANIN AND A. PALERMO

Figure 16. Pseudo-dynamic experimental test results for HBD3; top: cm1 record, bottom: lp5 record.

The simplest macro-model is a lumped plasticity model based on two rotational springs in
parallel at the rocking interface. Details of this model, along with the analytical–experimental
comparisons on beam–column joint subassemblies, wall systems, or bridge piers can be found in
[10, 16, 30, 32].
Presented herein is a macro-model based on multiple axial spring elements to replicate the
rocking interface (defined as a multi-spring element). In particular, multiple elastic, compression
only, spring elements represent the rocking interface between the pier and the foundation, whereas
axial spring elements model the behaviour of the unbonded post-tensioned tendons. Frame elements
are used to model the elastic cantilever behaviour of the pier, whereas non-linear axial springs
model the inelastic behaviour of the mild steel dissipaters as illustrated in Figure 17. The multi-
axial spring element [33] is implemented within the FE software Ruaumoko [34] and consists of 10
linear-elastic, compression-only, spring elements distributed along the length of the member. The
tendon spring elements are assigned stiffness to represent the actual unbonded length, geometry
and material properties of the 7-wire prestressing tendons. The behaviour of the external dissipater
elements are modelled as a non-linear inelastic springs adopting an appropriate steel hysteresis
rule [25] calibrated to component testing. The most challenging aspect of the multi-spring macro-
model is related to the calibration of the axial stiffness of the multi-spring contact element, where
a proper understanding is necessary to return reasonable results.

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
UNBONDED POST-TENSIONED ROCKING BRIDGE PIERS 349

Figure 17. Multi-(axial) spring macro-model.

In order to accurately determine the stiffness of the multi-spring unit, a moment-rotation section
analysis at the rocking interface is first defined. Unlike traditional moment-curvature analysis, strain
compatibility is violated at the rocking interface of a hybrid section due to: (a) an unbonded length
being provided for both the prestressed and non-prestressed reinforcement, and (b) a gap opening
mechanism resulting in infinite curvature. Based on a global member compatibility condition
equating global displacements between a hybrid section and an equivalently reinforced monolithic
section, a moment-rotation response can be generated [26, 30, 31].
Given the section response, the multi-spring stiffness can be calibrated. The process is a trial
and error procedure but given a good understanding and initial ‘guess’ of the axial stiffness,
the iteration process will converge rapidly. The calibration involves two parameters, namely the
neutral axis depth and the location to the centroid of the resultant compression force. The neutral
axis position will directly dictate the global stiffness of the connection via allocation of internal
member forces within both the prestressed and non-prestressed reinforcement. The location to the
centroid of the resultant compression force defines the internal lever arm of the section, and hence
the internal moment capacity of the section. It can be shown that in some situations the capacity
of the section can be relatively insensitive to the flexibility of the multi-spring interface (within
reasonable bounds), and thus a good match to the moment-rotation response can be obtained with
some approximation in the calibration. Furthermore, especially when using an elastic multi-spring
element, it is unlikely that a ‘perfect’ agreement between both the neutral axis depth and the
location to the centroid of the resultant compression force can be achieved, particularly when
considering a highly non-linear stress–strain behaviour at the interface as in the case of concrete to
concrete contact. An iterative method to compute the axial stiffness k is illustrated in Figure 18 via
a four-step process. Based on relatively limited calibration, an empirical equation is provided to
first estimate the axial stiffness, where Lcant is the cantilever length of the pier (in meters), B is the
width/depth of the square cross section (in meters) and E is the elastic modulus of pier material
(in GPa). The section moment M, neutral axis depth c/d and the centroid of the compression
region 0.5a are compared between the macro-model and the computed moment-rotation response.
A goodness of fit is computed for each relationship, based on the Coefficient of Determination R2 ,
or similar, and the axial stiffness is updated accordingly. Work is currently ongoing in this area to
provide numerical values to approximate the axial stiffness based on extensive calibration.

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
350 D. MARRIOTT, S. PAMPANIN AND A. PALERMO

Figure 18. Iterative method to determine the axial stiffness of a post-tensioned rocking interface.

For the model presented, a multi-spring unit with 10 uniformly distributed springs was used:
each spring having equal stiffness. A good agreement in calibration was found with an axial
stiffness of k = 5.5×106 kN/m.

COMPARISON WITH THE EXPERIMENTAL RESULTS

The initial validation of the macro-model was undertaken via a comparison to the quasi-static
testing (Figures 19 and 20) and then against the pseudo-dynamic testing (Figure 21). The ability
of the macro-model to capture the entire force–displacement relationship is well evident in both
Figures 19 and 20. The accuracy of the model stems from the ability to accurately model the
flexibility of the rocking interface. It follows that correctly modelling the rocking interface will
lead to a more accurate representation of the forces associated with each of the unbonded tendons
and mild steel dissipater elements. The model captures both the tendon elongation and rocking
interface well (as per the neutral axis depth versus lateral displacement).
The analytical–experimental comparison to the response of the hybrid specimen HBD4 presented
in Figure 20 further confirms the reliability of the model. The inaccuracy of the model to predict
the small residual displacements experienced during testing is mainly due to the inability of the
steel hysteresis rule [25] to accurately capture the observed increase in compressive load within
the dissipaters on unloading (as explained earlier and shown in Figure 9). However, the majority
of the inelastic dissipation, tendon load increment and neutral axis depth are captured well. The
macro-model also predicted failure of the steel dissipater elements during the final cycle of HBD4,
resulting in termination of the analysis.
Further validation of the analytical model is given by the comparison to the pseudo-dynamic
experimental testing (Figure 21). The tests revealed that the analytical model slightly overestimated
the initial stiffness of the system, resulting in some discrepancy within the time history domain. In
fact, while the overall behaviour is well captured, the maximum numerical displacement (model )
differs with respect to the experimental results by 15–20%. Further improvements in the model

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
UNBONDED POST-TENSIONED ROCKING BRIDGE PIERS 351

Figure 19. Numerical modelling of HBD3; left: force displacement, centre: tendon
behaviour, right: neutral axis behaviour.

Figure 20. Numerical modelling of HDB4; left: force displacement, centre: tendon
behaviour, right: neutral axis behaviour.

are currently being explored, including a non-linear multi-spring to capture a highly non-linear
rocking response.

CONCLUSIONS

The experimental response under quasi-static and pseudo-dynamic loading regimes of two hybrid
bridge piers with external replaceable dissipaters has been presented and compared with a traditional
monolithic cast in situ counterpart and a post-tensioned only solution.
The novelty of these high-performance seismic-resisting systems, based on unbonded post-
tensioning techniques, was the use of low cost, fully replaceable mild steel yielding dissipaters
fabricated ‘in-house’. Following a major earthquake, the dissipaters could be easily inspected
and replaced if necessary. Furthermore, the external nature of the dissipaters resulted in greater

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
352 D. MARRIOTT, S. PAMPANIN AND A. PALERMO

Figure 21. Experimental response and analytical modelling of HBD3; top:


cm1 record, bottom: lp5 record.

stability and energy dissipation when compared with more traditional hybrid technology consisting
of internally grouted mild steel reinforcement.
Testing of two hybrid bridge piers with replaceable, external mild steel dissipaters highlighted the
inherent stability, energy dissipation and re-centring characteristics when compared with traditional
reinforced cast in situ concrete. The design performance objectives consisting of allowable drift
and material strain limits were met based on quasi-static and pseudo-dynamic testing. Damage to
the monolithic pier consisted of wide flexural cracks, whereas the hybrid specimens were subjected
to minor flexural cracks which were no longer visible upon removal of the lateral load.
In general, the use of the proposed external, replaceable dissipaters has shown great promise for
use in the seismic protection of bridge structures. The quick erection time inherent of post-tension
construction techniques combined with inexpensive mild steel dissipaters (when compared with
more complex devices consisting of viscous, lead extrusion dampers or similar) makes this system
competitive with respect to traditional monolithic cast in situ bridge construction. Moreover, the
lack of damage within the structural members and the ease at which the dissipaters can be replaced,
drastically reduces the post-earthquake repair and economic cost.
Relatively simple macro-models based on a lumped plasticity approach (either rotational springs
in parallel or multi-spring contact elements) were adopted and used to further validate the

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
UNBONDED POST-TENSIONED ROCKING BRIDGE PIERS 353

experimental tests. Particular emphasis has been given to a refinement of the multi-spring model,
which if properly calibrated through a moment-rotation section analysis, provides good compar-
isons at both a local and global level.

ACKNOWLEDGEMENTS
The technical support from Mr Weng-Yuen Kam (PhD Candidate, University of Canterbury) and Mr
Michele Palmieri (visiting Laurea student, University of Brescia) during the pseudo-dynamic testing of
the specimens is greatly appreciated.

REFERENCES
1. Priestley MJN, Sritharan S, Conley JR, Pampanin S. Preliminary results and conclusions from the PRESSS
five-story precast concrete test building. PCI Journal 1999; 44(6):42–67.
2. Kurama YC. Hybrid post-tensioned precast concrete walls for use in seismic regions. PCI Journal 2002;
47(5):36–59.
3. Pampanin S. Emerging solutions for high seismic performance of precast/prestressed concrete buildings. Journal
of Advanced Concrete Technology 2005; 3(2):207–223.
4. Mander JB, Cheng CT. Seismic resistance of bridge piers based on damage avoidance design. Technical Report
NCEER-97-0014, State University of New York, Buffalo, NY, 1997.
5. Kwan W-P, Billington SL. Unbonded posttensioned concrete bridge piers. I: monotonic and cyclic analyses.
Journal of Bridge Engineering 2003; 8(2):92–101.
6. Palermo A, Pampanin S, Calvi GM. Concept and development of hybrid solutions for seismic resistant bridge
systems. Journal of Earthquake Engineering 2005; 9(6):899–921.
7. Kawashima K. Seismic design of concrete bridges. 1st fib Congress, Osaka, Japan, 2002.
8. Hewes JT, Priestley MJN. Experimental testing of unbonded post-tensioned precast concrete segmental bridge
columns. Proceedings of the 6th Caltrans Seismic Research Workshop Program, Sacramento, CA, 2001.
9. Ikeda S, Hirose S, Yamaguchi T, Nonaka S. Seismic performance of concrete piers prestressed in the critical
sections. 1st fib Congress, Osaka, Japan, 2002.
10. Palermo A, Pampanin S, Marriott D. Design, modelling and experimental response of seismic resistant bridge piers
with posttensioned dissipating connections. Journal of Structural Engineering (ASCE) 2007; 133(11):1648–1662.
11. Marriott D, Palermo A, Pampanin S. Quasi-static and pseudo-dynamic testing of damage resistant bridge piers with
hybrid connections. First European Conference on Earthquake Engineering and Seismology, Geneva, Switzerland,
2006.
12. Dhakal RP, Mander JB, Mashiko N. Identification of critical ground motions for seismic performance assessment
of structures. Earthquake Engineering and Structural Dynamics 2006; 35(8):989–1008.
13. Pinto A, Pegon P, Mangonette G, Tsionis G. Pseudo-dynamic testing of bridges using non-linear substructuring.
Earthquake Engineering and Structural Dynamics 2004; 33:1125–1146.
14. Cheok GS, Stone WC. Performance of 1/3 scale model precast concrete beam–column connections subjected to
cyclic inelastic loads. Report No 4: NISTIR 5436, National Institute of Standards and Technology, Gaithersburg,
1994.
15. Kurama YC, Pessiki S, Sause R, Lu L, El-Sheikh MT. Analytical modelling and lateral load behavior of unbonded
post-tensioned precast concrete walls. PRESSS Report No. 98/02, Lehigh University, 1998.
16. Marriott D, Pampanin S, Palermo A. Seismic design, experimental response, and numerical modeling of rocking
bridge piers with hybrid, post-tensioned connections. ISSN-0110-3326 (2007-01), University of Canterbury,
New Zealand, 2007.
17. Priestley MJN. Direct displacement-based design of precast/prestressed concrete buildings. PCI Journal 2002;
47(6):66–78.
18. ACI. Acceptance Criteria for Moment Frames Based on Structural Testing (T1.1-01) and Commentary (T1.1R-01).
Farmington Hills, MI, 2001.
19. Grant DN, Blandon CA, Priestley MJN. Modelling inelastic response in direct displacement-based design.
Research Report ROSE-2005/03, ROSE School, Pavia, 2005.
20. NZS1170.5. Structural Design Actions: Part 5 Earthquake Actions. Standards New Zealand Wellington, 2004.

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe
354 D. MARRIOTT, S. PAMPANIN AND A. PALERMO

21. Bracci JM, Reinhorn AM, Mander JB, Kunnath SK. Deterministic model for seismic damage evaluation of
reinforced concrete structures. NCEER Research Report NCEER-89-0033, State University of New York, Buffalo,
1989.
22. Kappos AJ. Seismic damage indices for RC buildings: evaluation of concepts and procedures. Progress in
Structural Engineering and Materials 1997; 1(1):78–87.
23. Park YJ, Ang AS. Mechanistic seismic damage model for reinforced concrete. Journal of Structural Engineering
(ASCE) 1985; 111(4):722–739.
24. Banon H, Veneziano D. Seismic safety of reinforced concrete members and structures. Earthquake Engineering
and Structural Dynamics 1982; 10:179–193.
25. Dodd LL, Restrepo-Posada JI. Model for predicting cyclic behavior of reinforcing steel. Journal of Structural
Engineering 1995; 121(3):433–444.
26. NZS3101. Concrete Structures Standard: Part 1—The Design of Concrete Structures. Standards New Zealand,
Wellington, 2006.
27. ACI. ACI Manual of Concrete Practice-Special Hybrid Moment Frames Composed of Discretely Jointed Precast
and Post-Tensioned Concrete Members: ACI T1.2-03. Farmington Hills, MI, 2007.
28. Marriott D, Pampanin S, Palermo A, Bull D. Dynamic testing of precast, post-tensioned rocking wall systems
with alternative dissipating solutions. Bulletin of NZSEE 2008; 41(2):90–103.
29. Priestley MJN, Calvi GM, Kowalsky MJ. Displacement-based Seismic Design of Structures. IUSS Press: Pavia,
2007.
30. Pampanin S, Priestley MJN, Sritharan S. Analytical modelling of the seismic behaviour of precast concrete
frames designed with ductile connections. Journal of Earthquake Engineering 2001; 5(3):329–367.
31. fib. Seismic Design of Precast Concrete Building Structures. fib Bulletin 27, Lausanne, 2003.
32. Palermo A, Pampanin S, Carr AJ. Efficiency of simplified alternative modelling approaches to predict the seismic
response of precast concrete hybrid systems. fib Symposium—Keep Concrete Attractive, Budapest, 2005.
33. Spieth HA, Carr AJ, Murahidy AG, Arnolds D, Davies M, Mander JB. Modelling of post-tensioned precast
reinforced concrete frame structures with rocking beam–column connections. 2004 NZSEE Conference, Rotarua,
2004.
34. Carr A. Ruaumoko2D. Inelastic Dynamic Time History Analysis Program, Christchurch, 2005.

Copyright q 2008 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:331–354
DOI: 10.1002/eqe

You might also like