Sie sind auf Seite 1von 246

SEISMIC PERFORMANCE AND FORENSIC ANALYSIS

OF A PRECAST CONCRETE
HOLLOW-CORE FLOOR SUPER-ASSEMBLAGE

A THESIS
SUBMITTED IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE

OF

MASTER OF ENGINEERING
AT THE

UNIVERSITY OF CANTERBURY
BY

CAMERON MACPHERSON

UNIVERSITY OF CANTERBURY
CHRISTCHURCH, NEW ZEALAND
2005
TABLE OF CONTENTS

Abstract...vi
Acknowledgements... vii

1 INTRODUCTION...1-1
1.1 Introduction..1-1
1.2 Previous Relevant Research 1-3
1.3 Findings from Matthews (2004)...1-4
1.3.1 Origin of the Test Specimen Super-Assemblage. 1-4
1.3.2 Loading System... 1-5
1.3.3 Major Findings: Beam-to-Floor Slab connection 1-6
1.3.4 Major Findings: Lateral Frame-to-Floor connection... 1-7
1.3.5 Analytical Findings.. 1-8
1.4 Findings from Lindsay (2004) 1-9
1.4.1 Hollow-core Connection Details. 1-9
1.4.2 Major Conclusions and Recommendations... 1-11
1.5 Outline of this Thesis. 1-13
1.6 What is new and significant in this Thesis?...1-15
1.7 References..1-16

2 SUB-ASSEMBLAGE EXPERIMENTAL INVESTIGATION 2-1


Section Summary..2-1
2.1 Introduction..2-2
2.2 Previous Research Findings and New Zealand Guidelines. 2-3
2.3 Test Specimen Details... 2-11
2.3.1 Specimen 1.2-11
2.3.2 Specimen 2.2-12
2.4 Beam Elongation... 2-13
2.4.1 Sub-Assemblage Elongation.. 2-15
2.5 Experimental Set-up...2-17
2.6 Experimental Results: Specimen 1 2-20
2.7 Experimental Results: Specimen 2 2-23

Page i
2.8 Strength and Failure Mode Prediction Models.. 2-28
2.8.1 Moment Capacity...2-28
2.8.2 Low Cycle Fatigue. 2-29
2.9 Discussion of Results. 2-30
2.10 Conclusions2-32
2.11 References..2-32

3 SUPER-ASSEMBLAGE EXPERIMENTAL INVESTIGATION 3-1


Section Summary. 3-1
3.1 Introduction..3-2
3.2 Findings from Previous Research 3-3
3.3 Super-Assemblage Specimen Design Details and Construction. 3-6
3.3.1 Hollow-core Seating Details 3-6
3.3.2 Lateral Connection and Diaphragm Reinforcement.3-8
3.3.3 Super-Assemblage Construction. 3-9
3.3.4 Material Properties.3-11
3.4 Experimental Set-up...3-12
3.4.1 Instrumentation.. 3-12
3.4.2 Loading Protocol... 3-16
3.5 Experimental Observations 3-17
3.5.1 Phase 1: Longitudinal Loading 3-17
3.5.2 Phase 2: Transverse Loading. 3-19
3.5.3 Phase 3: Longitudinal Re-Loading 3-20
3.6 Hysteretic Performance and System Strengths.. 3-23
3.7 Discussion of Results and Effectiveness of Design Changes 3-25
3.7.1 Transverse Beam Torsion.. 3-25
3.7.2 Hollow-core-to-Supporting Beam Connection.. 3-29
3.7.3 Supporting Beam Detailing Enhancements... 3-30
3.7.4 Moment Timber Infill Connection and Diaphragm Reinforcing... 3-31
3.7.5 Displacement Incompatibility 3-32
3.8 Conclusions3-34
3.9 References..3-35

Page ii
4 FORENSIC ANALYSIS OF FAILURE AND BEHAVIOUR MODES.. 4-1
Section Summary. 4-1
4.1 Introduction..4-2
4.2 Torsional Behaviour of the Transverse Support Beams.. 4-4
4.2.1 Determination of the Torsional Demand. 4-9
4.2.2 Transverse Beam Torsional Capacity 4-12
4.3 Displacement Incompatibility 4-14
4.3.1 Identified Issues and Previous Findings 4-14
4.3.2 Analytical Modelling of Displacement Incompatibility 4-17
4.3.3 Scope for Future Advancements 4-19
4.4 Lateral Load Capacity of the System.4-20
4.4.1 Bowstring Effect 4-21
4.4.2 General System Strength... 4-26
4.4.3 Localised Strengths 4-28
4.5 Beam Elongation... 4-31
4.6 Low Cycle Fatigue of the Reinforcement.. 4-33
4.7 Comparisons with Previous Research... 4-34
4.7.1 Performance of the Seated Connection..4-35
4.7.2 Performance of the Perimeter Frame to Hollow-Core Connection... 4-37
4.7.3 Overall Performance Comparisons 4-38
4.8 Conclusions4-40
4.9 References..4-41

5 CONCLUSIONS AND RECOMMENDATIONS... 5-1


5.1 Summary.. 5-1
5.2 Conclusions..5-2
5.3 Recommendations for Professional Practice... 5-5
5.4 Recommendations for Future Research... 5-7
5.5 References5-9

Page iii
APPENDIX A SUB-ASSEMBLAGE EXPERIMENTS: CONSTRUCTION & RESULTS...A-1
A.1 Construction Details and Photographs A-1
A.2 Material Tests..A-5
A.3 Testing Photographic Log... A-1
A.3.1 Experiment Set-up and Instrumentation. A-6
A.3.2 Specimen 1.. A-7
A.3.3 Specimen 2.. A-8
A.4 Loading Protocol.......................A-10
A.5 Low Cycle Fatigue Prediction...A-11
A.6 References. A-12

APPENDIX B SUPER-ASSEMBLAGE CONSTRUCTION PHOTOGRAPHS AND DESIGN


DETAILS.B-1
B.1 Building Construction Drawings.B-1
B.2 Building Construction Photographic Log. B-20

APPENDIX C SUPER-ASSEMBLAGE TESTING AND PHOTOGRAPHIC LOG.. C-1


C.1 Material Testing...C-1
C.1.1 Reinforcement Testing C-1
C.1.2 Concrete Testing. C-3
C.1.3 Hollow-core Properties... C-4
C.2 Instrumentation Photographs.. C-5
C.3 Testing Photographic Log... C-7
C.3.1 Phase 1: Longitudinal loading.C-7
C.3.2 Phase 2: Transverse loading... C-9
C.3.3 Phase 3: Longitudinal re-loading C-10
C.3.4 End of Testing... C-13
C.4 Concrete Topping Delamination... C-14

APPENDIX D ANALYTICAL RESULTS... D-1


D.1 Yield Drift CalculationD-1
D.1.1 Beam Contribution.. D-1
D.1.2 Beam Column Joint Contribution... D-2

Page iv
D.1.3 Column Contribution.. D-2
D.2 Super-Assemblage StiffnessD-4
D.3 Torsion Test D-5
D.4 General Super-Assemblage Performance... D-6
D.4.1 Central Column Joint Reinforcing Bar Slip D-6
D.4.2 Hollow-core floor unit pull-off... D-8
D.4.3 Beam Column Joint Rotation Contributions.. D-10
D.4.4 Reinforcement Strain Information.D-12
D.5 West Transverse Beam Torsion D-18
D.5.1 Experimental Results D-18
D.5.2 Calculation of Torsional Demand. D-19
D.5.3 Calculation of Torsional Capacity D-20
D.6 Displacement Incompatibility... D-23
D.6.1 Experimental Results D-23
D.6.2 Analytical Modelling D-28
D.7 Beam Elongation...D-31
D.8 Low Cycle Fatigue Fracture of Reinforcement D-34
D.9 Observed Bowstring Effect... D-36
D.10 Super-Assemblage Capacity Mechanism Calculations D-38
D.10.1 Phase 1 and 3: Longitudinal ... D-38
D.10.2 Phase 2: Transverse.. D-40
D.11 References. D-46

Page v
ABSTRACT

Recent earthquakes and laboratory research has demonstrated the seismic vulnerability of
buildings constructed with precast prestressed concrete hollow-core floor systems. However,
further investigations have shown that with simple detailing enhancements, significant
improvement in the seismic performance of hollow-core floor systems can be expected. The
New Zealand Concrete Structures Standard provides two acceptable solutions for the
connection of hollow-core floor units to concrete supporting beams (Amendments to NZS
3101, 2004). The second of these acceptable solutions is currently untested, and the present
experimental investigation aims at validating this solution as well as several other new
detailing enhancements. Two dimensional sub-assemblage component experiments are
conducted utilising a newly developed, more realistic loading configuration. Based on the
sub-assemblage research findings, a concrete frame super-assemblage is constructed
incorporating a reinforced connection that rigidly ties the floor into the supporting beam and
an articulated topping slab cast onto a timber infill solution that bridges between the hollow-
core units and parallel longitudinal frame beams. The full-scale structure is cyclically tested
in both the longitudinal and transverse directions to inter-storey drifts of 5%. Visual and
instrumental observations from the experiment are presented and discussed. A forensic
analysis of the experimental observations is performed. Torsional hinging behaviour that is
observed due to a strong floor-to-beam connection and weak support beam is explained and
experimentally observed displacements are compared with computational modelling. The
lateral load capacity of the super-assemblage is predicted and it is seen that it is necessary to
incorporate all sources of strength enhancement. Finally, the results from this present
investigation are compared and contrasted with similar previous research investigations and
recommendations for the forthcoming revision of the New Zealand Concrete Standard, NZS
3101, are made. The experimental observations demonstrate that with appropriate detailing
enhancements there can be significant seismic behaviour improvements whereby a relatively
fragile hollow-core flooring system is transformed into a robust frame-floor system where all
the damage is transferred into the plastic hinge zones of the supporting moment resisting
frame.

Page vi
ACKNOWLEDGEMENTS

This research was undertaken in the Department of Civil Engineering, at the University of
Canterbury under the supervision of Professor John Mander and Professor Des Bull. I would
like to thank John and Des for their support, encouragement, expertise and guidance
throughout this research and my university studies.
This project would simply not have been possible without the support of many
companies and organizations. Financial support has been made for this research by a grant
from the EQC Research Foundation. Products and support were provided by the following
organisations: University of Canterbury, Firth Industries Ltd, Stresscrete, Pacific Steel and
Fletcher Reinforcing, Construction Techniques and McDowels Concrete Accessories. All of
this financial and in-kind support is gratefully acknowledged.
Jeff Matthews and Renee Lindsay, from whom this research has continued on from,
deserve enormous amounts of thanks. Their friendliness, knowledge and experience from
their studies have helped me considerably. Jeff and Renees high standard of work has also set
high qualities for me to aspire to.
I would like to thank John Trowsdale for his support throughout my research, and
acknowledge his work on the sub-assemblage part of this project. Johns practical expertise
and friendliness were invaluable in completing the experiments. I would also like to thank my
fellow postgraduate students for their friendship and understanding of postgraduate life.
Full credit must go to all of the technical staff of the Civil Engineering Department. I
am thankful for everyones helpfulness and guidance. I would like to give special thanks to
Russel McConchie, Nigel Dixon, Tim Perigo and Richard Newton for the amount of effort
they have put into my research and for all of their support and friendliness that made my
experimental work so enjoyable. I would also like to thank Dr Stefano Pampanin for his
guidance, expertise and encouragement. I have learnt a lot from Stefano throughout my
postgraduate studies.
Finally, I would like to thank my friends and family for all the love and support and
most especially, Ryoko for her guidance, patience and support throughout my time at
university.

Page vii
1. Section 1 Introduction

1.1 INTRODUCTION

Recently, precast concrete construction has become the conventional form of construction in

New Zealand. The ease of construction, rapid erection times and the appeal of the off-site

construction and factory-like controls available in the precast, prestressed industry are reasons

for its popularity. Since the 1980s, one of the most common construction types in New

Zealand has been the use of precast prestressed hollow-core floor units seated on reinforced

concrete moment resisting frames. However, the proliferation of precast concrete construction

has preceded research in many areas and there are concerns over the seismic performance of

these hollow-core floor systems in New Zealand. The possibility for the loss of seating to

occur, lack of composite bond strength with the topping concrete, cracking of the webs of the

hollow-core units and the general lack of structural redundancy have been highlighted as main

factors that could lead to poor performance and collapse of several precast concrete buildings.

Over the last few years, research at the Universities of Auckland and Canterbury has sought to

investigate the seismic performance of these precast systems and the level of seismic

vulnerability of New Zealand buildings.

Recently, research was conducted to fully assess the adequacy of a typical New

Zealand designed and built concrete structure with hollow-core floors (Matthews, 2004;

Matthews et al, 2003, 2004). A full-scale three-dimensional section of a common New

Zealand multi-storey concrete frame building with hollow-core floor units was constructed.

Traditionally, research effort has either focused at a structural component level or at

investigating lateral load resisting systems, and the effects of floor diaphragms have often

been overlooked by the simplification of the rigid diaphragm assumption. The results from the

Page 1-1
Matthews (2004) experiment showed that precast concrete structures in New Zealand may not

perform as well as would be hoped and the collapse of floors could be expected. The overall

collapse of the floor units during the experiment showed similar traits to some failures

observed after the Northridge (1994) earthquake (Norton et al, 1994), and the main concern

became the detailing used for the connections of the hollow-core floors to the concrete frame.

The investigation into the seismic vulnerability of existing buildings provided several

questions requiring immediate answering. In particular, new solutions needed to be obtained

for future construction, the retrofit of existing structures with hollow-core floors, as well as

further investigations into the problems. The issues surrounding the use of prestressed floors

that have stemmed from the Matthews (2004) experiment, previous research and earthquake

events, have been recognised in New Zealand Concrete Design Standards. Specific

requirements are in place prescribing minimum seat widths and different connection solutions

that will address the shortcomings highlighted with traditional construction methods

however it is important to verify design standard details analytically and experimentally. In

response to the Matthews (2004) testing a repair was carried out on the concrete super-

assemblage and second phase testing was undertaken by Lindsay (2004) and Lindsay et al

(2004). This experiment investigated improved connection details for the hollow-core floor

systems for new construction that have been included in proposed amendments to the New

Zealand Concrete Design Standard (NZS 3101, 2004).

This thesis follows the research completed by Matthews (2004) and Lindsay (2004).

This chapter initially briefly discusses the previous sub-assemblage and super-assemblage

research related to hollow-core floor systems. A summary of the Matthews (2004) and

Lindsay (2004) experiments is provided, including the origins of the test specimen, the

development of the loading system, and a review of the main test findings. The main sections

of the thesis are then outlined.

Page 1-2
1.2 PREVIOUS RELEVANT RESEARCH

The performance of hollow-core floor systems has been examined in many experimental

programs at a two-dimensional sub-assemblage level. Meija-McMaster and Park (1994),

Herlihy et al (1995), Oliver (1998) and Herlihy and Park (2000) have all examined the seat

requirements of hollow-core floor units due to beam elongation for a variety of end

connection details. Further research by Bull and Matthews (2003) and Liew (2004) have

similarly tested various connection details, but introduced the relative rotation between the

floor slab and supporting beam as the main cause of damage. A more in-depth summary of

previous research is located in Section 2 of this thesis.

There have been many super-assemblage research programs relevant to this project

involving concrete frame structures with and without floor slabs. Zerbe and Durrani (1989,

1990), Fenwick and Megget (1993), Restrepo (1993), Fenwick et al (1995) and Lau (2002)

have conducted tests examining beam elongation and the contribution of floor slabs to the

overall performance and strength of lateral load resisting elements. However, due to the

respective test set-ups used, beam elongation has been either promoted or restrained and the

results were not realistic. Despite sub-assemblage research previously investigating hollow-

core seating connections, prior to the Matthews (2004) investigation, a full-scale experiment

examining the seismic adequacy of hollow-core floor systems had not been undertaken. The

need for these tests is due to three-dimensional effects, such as beam elongation, torsion and

displacement incompatibility between the perimeter concrete frames and flooring, which all

must be present to ensure realistic performance is represented. The particular issue of

displacement incompatibility has been acknowledged as a source of damage (Lau et al, 2002

and Priestley et al, 1999), but little quantitative research has been done to explore the issue.

Page 1-3
1.3 FINDINGS FROM MATTHEWS (2004)

Owing to the widespread use of hollow-core flooring in precast concrete construction, an

experimental program was initiated to examine the seismic adequacy of such systems in the

New Zealand building stock. A two bay by one bay concrete moment resisting frame

performed relatively well whereas the hollow-core flooring system behaved poorly and

collapsed at moderate levels of seismic intensity. The following subsections briefly describe

the origin and details of the super-assemblage including the seated and lateral connections of

the hollow-core floor units, the development of the self-equilibrating load frame and the main

observations and analytical results from the experiment.

1.3.1 Origin of the test specimen super-assemblage

The super-assemblage is a one-storey corner segment of a typical multi-storey precast

concrete moment resisting frame building built in New Zealand over the last two decades.

Figure 1-1 shows a view of a representative structure from which test specimen was derived

from. The corner section was chosen to examine the effects of earthquake actions in different

directions - longitudinally as well as orthogonal to the span of the flooring system. The bay

widths were 6.1m long and the 300 series hollow-core flooring spans over two bays past an

intermediate column and is seated on the two end beams. Although this type of construction

whereby the double span flooring is present is not the most common form, it was chosen to

investigate a worst-case scenario. Buildings of this type have been constructed in Wellington,

which is a high seismicity region of New Zealand, whereas the use of double span hollow-

core is less frequent in Auckland and other regions of New Zealand. For these reasons, the

layout of super-assemblage has been questioned (OGrady, 2004). As the following

descriptions of the findings from the Matthews (2004) test will indicate, the serious

Page 1-4
inadequacies of the hollow-core flooring were centred on the supporting beam-to-slab

connection, which is irrespective of the hollow-core span.

Selected portion to be tested

Figure 1-1 Selection portion representing the test super-assemblage (Matthews, 2004)

1.3.2 Loading System

An integral part of the Matthews (2004) experimental program was the development of the

loading frames. As a strong wall or strong floor was unavailable to provide reaction support

for loading, a self-equilibrating load frame was constructed to load the structural super-

assemblage. The loading frames were designed and implemented such that under lateral

loading, the building deforms in a realistic manner and that the advent of beam elongation is

neither restrained nor promoted. The self-equilibrating primary loading frame is shown in

Figure 1-2. By controlling the actuators within the loading frame, equal and opposite diagonal

loads are applied to the tops and bottoms of the columns ensuring that the system is in

equilibrium (no axial forces are induced in the beams) and that axial forces cancel within the

system.

To ensure that the columns displace parallel to each other as they would in a real

building, a secondary load frame (SLF) was used. The frame, as shown in Figure 1-2, acts to

keep the columns parallel to each other but also allows beam elongation to occur through a

Page 1-5
reduction of the initial angle of the frame. The design of the loading frames and test specimen

allows the super-assemblage to be tested in both longitudinal and transverse directions. For

longitudinal loading, additional rams are attached at the base of the back columns to maintain

the same column inclinations and ensure that any artificial torsion in the transverse beams

does not occur.


Secondary
loading frame Secondary
loading frames

Ac B
tua tor
tor tua
D Ac

E Ac
tor tu a
tua tor
Ac A
Primary
Loading frame Primary
Loading frames

Side Elevation Front Elevation

Figure 1-2 Self equilibrating and secondary loading frames used by Matthews (2004)

1.3.3 Major Findings: Beam-to-Floor Slab Connection

In compliance with the overlying investigation into the seismic abilities of existing buildings,

the test super-assemblage included the floor slab connection with the supporting beam

predominantly used in New Zealand construction practice. The seating connection featured a

300 series hollow-core unit seated on a mortar bed on either a 20mm or 40mm seating. The

cores at the end of the floor units were plugged with common plastic end bungs to prevent

concrete from entering the cores. A 75mm thick topping slab was cast on top of the floor units

and conventional Grade 300 starter bar reinforcement was used together with the cold drawn

mesh in the topping slab. Initial damage to the supporting beam-to-floor slab connection

during testing occurred earlier than expected and the mode of failure was different to what

was previously assumed. Originally it was thought that loss of seating occurred due to the

floor units sliding off the seat of the supporting beams. However, it became apparent that the

Page 1-6
relative rotation between the flooring and beam was the major source of damage. The mortar

bed possessed sufficient rigidity to ensure that the floor units could not slide, and the negative

moments induced by the relative rotation caused diagonal compressive struts and orthogonal

tensile stresses, led to splitting of the unreinforced webs of the hollow-core. Positive moments

created large crack openings, and under cyclic loading the connection deteriorated

considerably, ultimately causing the entire support at both ends of the structure to be lost. A

photograph illustrating the overall failure of the flooring is shown in Figure 1-3, and a

diagram of the supporting beam floor slab connection is shown in Figure 1-4(a).

1.3.4 Major Findings: Lateral Frame-to-Floor Connection

Figure 1-4(b) shows the connection used for the Matthews (2004) experiment between the

perimeter moment resisting concrete frame and the adjoining parallel hollow-core unit. In

addition to the damage incurred by the end connection, the first hollow-core unit adjacent to

the perimeter frame showed signs of distress at an early stage of testing. The construction

connection inadvertently rigidly tied the floor into the perimeter beams, and under lateral

displacements, the flooring was forced to displace in the double curvature pattern of the frame

beams. Hollow-core flooring sags by nature and this undesirable deformation caused internal

cracking of the webs of the hollow-core unit and the eventual failure of the lower half of the

first unit.

The displacement incompatibility coupled with the tension forces in the floor

diaphragm caused the cold drawn mesh reinforcement within the topping slab to fracture and

a large tear to form within the topping concrete along the interface between the first and

second hollow-core units spanning parallel to the perimeter frame. This caused the centre

column of the front frame to displace outwards, transverse to the plane of loading. The

broader ramifications of this can be seen if this outward column movement was applied to

Page 1-7
several floors of a multi-storey structure during an earthquake. Premature buckling of the

interior intermediate columns could result in the potential collapse of the structure. This also

identifies the necessity to tie the intermediate column into the floor diaphragm, and this was

recommended for further research.

Figure 1-3 Overall failure of the floor units in the Matthews (2004) experiment

1.3.5 Analytical Findings

In addition to the experimental investigation conducted by Matthews (2004), significant

analytical work was undertaken to complement the observed results. A capacity mechanism

analysis was used to calculate the overall strength capacity of the super-assemblage. A

method was proposed to determine the contribution of the floor diaphragm to the overall

performance. Similarly, a theory was developed by Matthews (2004) and Matthews et al

(2004) to predict beam elongation. The method uses rigid body kinematics and a rainflow

counting technique to predict growth of plastic hinges under increasing rotations. For both

beam elongation and the strength capacity prediction models, the comparisons with the

experimental results were good.

Page 1-8
1.4 FINDINGS FROM LINDSAY (2004)

A continuation of the work done by Matthews (2004) was conducted by Lindsay (2004). This

investigation aimed to verify recommended details for the improved performance of future

precast concrete construction practice, which is in contrast to Matthews (2004) who

investigated past construction techniques (1985-2000) and the resulting problems with

seismic performance. Lindsay (2004) repaired and reconstructed the Matthews (2004) full-

scale super-assemblage specimen and then retested it under a more rigorous loading regime.

Modifications were made to improve the performance of the hollow-core floor system for

both the lateral and end seating connections between the frame and hollow-core units, as well

as more adequately tying the central column into the floor diaphragm.

1.4.1 Hollow-core Connection Details

As shown in Figure 1-4(c), Lindsay (2004) employed a beam to floor connection detail that

used a low-friction bearing strip and a compressible backing board, which replaced the plastic

end plugs in the cores at the end of the units. The seating connection detail used during the

Matthews (2004) test could be likened to a semi-rigid type of connection. The poor

behaviour led to further testing of a pinned type of connection by Lindsay (2004), whereby

the low friction bearing strip allows the floor units to slide horizontally relative to the

supporting beam, and the compressible backing board to compress reducing large

compression stresses entering the floor unit. The overall performance was satisfactory, and

significant improvements were seen relative to the Matthews (2004) support condition. The

connection was able to sustain sizeable relative rotations with overall structural inter-storey

drifts reaching +5%. The compressible backing board did not compress as expected which

indicates beam elongation effects were causing pull-off of the flooring units off the seating

Page 1-9
away from the supporting beam. The low friction bearing strips also did not perform as

originally expected, whereupon the rough side of the bearing strip slid on the concrete seat

and the low friction side adhered to the floor unit. Significant spalling of the concrete ledge

(seat) of the supporting beams later in the experiment reduced the seat width to minimal

amounts and further jeopardised the integrity of the support connection.

Zone of Grade 430 starters


Conventional Grade 300 starters at 300 crs weakness at 300 crs & 665 mesh
end-plug & ductile mesh in topping
in 75mm topping

75mm concrete
topping

Mortar bed 300 series


hollow-core unit First
Second Hollow-core
20mm seat East end Hollow-core Unit
40mm seat West end Unit
Perimeter beam
Column face
(a) Matthews (2004) floor-slab connection (b) Matthews (2004) lateral connection
Zone of weakness
5mm Compressible Grade 300 starters at 300 crs
Timber in-fill
material & ductile mesh in topping Existing grade 430 starters at 300 crs
750mm
& ductile mesh in 75mm topping

Low Friction Hollow-core Units


bearing strip Perimeter beam

75mm seat Column face

(c) Lindsay (2004) floor-slab connection (d) Lindsay (2004) lateral connection

Figure 1-4 Hollow-core connection details used by Matthews (2004) and Lindsay (2004)

To overcome the issues of the displacement incompatibility that occurs between the perimeter

frame and the floor units, a topping slab was cast onto a timber infill that bridged between the

hollow-core units and perimeter frame beams. Figure 1-4(d) shows a diagram of the

connection. The timber infill was proposed subsequent to the poor behaviour witnessed in the

Matthews (2004) experiment to allow more desirable one-way hollow-core action. The thin

slab was designed to provide a flexible interface and accommodate the displacement

Page 1-10
incompatibility between the floor and perimeter beams. Overall, the infill detail showed a

marked improvement (when compared to Matthews (2004). However the ductile mesh used to

reinforce the topping slab fractured and a major longitudinal tear formed along the interface

between the infill and hollow-core, which interfered with the transfer of inertia and structural

interaction forces from the floor to the moment resisting frame. The presence of the drag bar

reinforcement tying the centre column into the floor plate ensured that there was no outward

translation of the centre column similar to what was observed during the Matthews (2004

experiment.

1.4.2 Major Conclusions and Recommendations

The improved detailing utilised in terms of the seated and lateral hollow-core connection

details and adequate tying of the central column into the floor, meant that the super-

assemblage was able to sustain inter-storey drifts up to +5%, until the low cycle fatigue of the

beam longitudinal reinforcement occurred. Although the experiment was success in many

areas, it also uncovered other issues that the experiment was unable to address. Some of the

detailing alterations made from the Matthews (2004) experiment still required further

enhancement and the following recommendations for professional practice and improving

design standards were made by Lindsay (2004):

1. Cold drawn or ductile mesh should not be used and conventional reinforcement should

be used in its place for diaphragm topping reinforcement over precast hollow-core

units.

2. In the presence of an infill slab offsetting the flooring from adjacent frame beams, the

starter bars used should extend over the infill, across the infill-to-hollow-core floor

interface and extend above the first floor slab unit. This will increase the ductility at the

infill-floor slab interface and minimise the risk of reinforcement fracture.

Page 1-11
3. Hollow-core units should not be seated on potential plastic hinge zones of supporting

beams. This should be overcome by having an infill slab over the highly deformable

hinge region, and hooked longitudinal reinforcement within the supporting beam to

concentrate large deformations away from the seated hollow-core units and close to the

column face.

4. A second generation low friction bearing strip should be employed for future use,

whereby the floor slab slides horizontally on the strip relative to the supporting beam,

and the strip adheres to the concrete seating. A strip featuring longer teeth on the

rough side will help to minimise spalling by evening out the surface roughness of the

seat concrete and help to ensure that the bearing strip will stay on the seat as designed.

In addition, transverse reinforcement should be used within the concrete seating of the

supporting beams.

Similarly, several recommendations were made by Lindsay (2004) for future research. These

include the following:

1. Due to the large amount of time and resources required to undertake large full-scale

super-assemblage tests, more two-dimensional proof-of-concept tests should be

performed, similar to those performed by Bull and Matthews (2003) and Liew (2004).

Despite the limitations of this method of testing, the comparable performance and

failures observed with full-scale three-dimensional experiments mean that sub-

assemblage tests are of great use in order to easily and relatively quickly assess the

adequacy of many other types of seated connections. Such tests could allow the

generation of a database of results and allow for a wide range of details to be tested

including all the different types of precast elements used in New Zealand construction.

The effects of beam elongation should also be incorporated into the testing procedure.

Page 1-12
2. Of a more urgent nature, the second of the two details included in the 2003 Amendment

to the New Zealand Concrete Structures Standard (NZS 3101, 2004) needs to be tested

as part of a full-scale three-dimensional super-assemblage experiment.

3. Owing to the inadequacies associated with hollow-core flooring connection details in

existing buildings, a retrofit connection solution needs to be developed and tested such

that life-safety in existing buildings can be maintained.

Other questions relating to hollow-core seated connection details were raised including; what

is the effect of negative seating (a bridged connection detail when the hollow-core units are

cut too short)? What is the effect of different beam geometries? What are the effects of

placing additional shear reinforcement in the cores of the units? Recommendations were also

made for further analytical research. The torsion of the transverse beams needs to be modelled

to gain an idea of when full activation of the starter bars occurs. Also analytical modelling of

the diaphragm, and in particular, the bowstring effect needs to be conducted.

1.5 OUTLINE OF THIS THESIS

This thesis comprises of three main sections. Following this introductory section, Section 2

discusses the testing of two sub-assemblage beam-to-hollow-core specimens. Previous

research at a sub-assemblage level is more closely reviewed and the experimental set-up is

described. The results are presented and from which the conclusions and recommendations

are drawn.

Section 3 covers the construction of the moment resisting frame and the modifications

made from the original Matthews frame (2004), as well as the final super-assemblage

specimen details and hollow-core connection details. The experimental loading regime and

instrumentation are outlined. The second part of this section provides visual and instrumental

Page 1-13
observations from the testing, hysteretic performance of the structure and a discussion of the

major results.

The penultimate section, Section 4, provides a forensic analysis of the experimental

observations. Modelling of displacement incompatibility between the lateral load-resisting

frame and the first floor slab unit is undertaken and comparisons between the analytical

results and experimental observations are made. The lateral system capacity is calculated from

the individual components of the system and compared with the experimental results. The

bowstring effect of floor slabs is also examined. The beam elongation results from the

experiment will be compared to theoretical predictions. Finally, a comparison will be made

between this experiment, Matthews (2004) and Lindsay (2004) in terms of the floor seating

connection, the lateral connection between the frame and initial hollow-core unit, and the

overall system strengths.

The final section of this thesis provides a summary and major conclusions from the

research presented. Recommendations are made for professional practice and future research.

Considerable material has also been included at the end of this thesis in the various

appendices. This material supplements the information provided in the main sections of the

thesis.

Appendix A consists of additional material from the sub-assemblage experimental

investigation and features photographic logs of the construction and testing as well as

extra construction details.

Appendix B comprises of a construction drawings and a photographic log of the

construction of the main super-assemblage structure.

Appendix C provides supplementary material to the testing of the super-assemblage.

Details and photographs of the test configuration are included as well as a

photographic record of the testing.

Page 1-14
Appendix D offers additional instrumental results from the testing and calculations to

supplement the various theoretical work undertaken throughout this thesis.

1.6 WHAT IS NEW AND SIGNIFICANT IN THIS THESIS?

The investigation presented in this thesis experimentally examines the seismic capabilities of

a reinforced hollow-core support connection detail that has been proposed in amendments to

the New Zealand Concrete Design Standard (NZS 3101, 2004). To this point, the specific

detail has not been tested, and this research describes the performance of the connection detail

in a new and improved two-dimensional experiment configuration before more realistic

testing in a three-dimensional full-scale structure. A precast concrete super-assemblage

specimen has been constructed and incorporates the new detailing enhancements that seek to

address the shortcomings of hollow-core floor systems that have been highlighted from

previous research.

Analytical work has been conducted to aid the understanding of displacement

incompatibility between the floor and frame elements of a structure. Non-linear modelling has

been undertaken and the results compared with actual values obtained from the experiment.

The model has then been extended and examples of the possible applications of the model are

discussed. The bowstring effect in floor slabs is examined and a method is introduced to allow

for the influence it has on the lateral load capacity of the super-assemblage.

The overall findings from this research produce a greater level of understanding of

hollow-core floor systems and the three-dimensional interaction between floor slabs and

reinforced concrete frames. Important recommendations are made for both professional

design practice and construction as well as providing ideas for future research. According to

the results of the hollow-core connection detailing, direct recommendations are made for the

upcoming revised New Zealand Concrete Design Standard.

Page 1-15
1.7 REFERENCES

Bull D.K, Matthews J.G, 2003, Proof of Concept Tests for Hollow-core Floor Unit

Connections, Precast NZ report, Feb 2003.

Fenwick R.C, and Megget L.M, 1993, Elongation and load deflection characteristics of

reinforced concrete members containing plastic hinges, Bulletin of NZNSEE, Vol 26,

No. 1, March, pp28-41.

Fenwick R.C, Davidson B.J, and McBride A, 1995, The influence of slabs on elongation in

ductile seismic resistant concrete frames, Proceedings for NZNSEE technical

conference, Rotorua, March, pp36-43.

Herlihy M.D, Park R and Bull D.K, 1995, Precast Concrete Floor Unit Support and

Continuity, Research Report 93/103, Department of Civil Engineering, University of

Canterbury, Christchurch, New Zealand.

Herlihy M.D, and Park R 2000, Precast concrete floor support and diaphragm action,

Research Report 2000/13, Department of Civil Engineering, University of Canterbury,

Christchurch, New Zealand.

Lau D.B.N, Fenwick R.C and Davidson B.J, 2002, Seismic performance of R/C perimeter

frames with slabs containing prestressed units, Proceedings for NZSEE conference,

Napier, March, Paper No. 7.2.

Lindsay R.A, 2004, Experiments on the seismic performance of hollow-core floor systems in

precast concrete building, Masters Thesis, Department of Civil Engineering,

University of Canterbury, Christchurch, New Zealand.

Lindsay R.A, Mander J.B, and Bull D.K, 2004a, Preliminary results from experiments on

hollow-core floor systems in precast concrete buildings, 2004 NZSEE Conference,

Rotorua, NZ, Paper 33, 8 pp.

Page 1-16
Lindsay R.A, Mander J.B, and Bull D.K, 2004b, Experiments of the Seismic Performance of

Hollow-core Floor Systems in Precast concrete Buildings, 13th World conference on

Earthquake Engineering, Vancouver, Canada, CD ROM paper No. 585.

Liew H.Y, 2004, Performance of Hollow-core Floor Seating Connection Details, Masters

Thesis, Department of Civil Engineering, University of Canterbury, Christchurch,

New Zealand.

Matthews J.G, 2004, Hollow-core floor slab performance following a severe earthquake, PhD

Thesis, Department of Civil Engineering, University of Canterbury, Christchurch,

New Zealand.

Matthews J.G, Bull D.K, and Mander J.B, 2003a, Background to the Testing of a Precast

Concrete Hollowcore Floor Slab Building, 2003, Pacific Conference on Earthquake

Engineering, February, Christchurch, NZ, CD ROM paper 170.

Matthews J.G, Bull D.K, and Mander J.B, 2003b, Preliminary Experimental Results and

Seismic Performance Implications of Precast Floor Systems with Detailing and Load

Path Deficiencies, 2003 Pacific Conference on Earthquake Engineering, February,

Christchurch, NZ. CD ROM paper 077.

Matthews J.G, Mander J.B, Bull D.K, 2004, Prediction of beam elongation in structural

concrete members using a rainflow method, 2004 NZSEE Conference, Rotorua, NZ,

Paper 27, 8 pp.

Meija-McMaster J.C and Park R, 1994, Tests on Special Reinforcement for the End Support

of Hollow-core Slabs, PCI Journal, Vol 39, No. 5, pp. 90-105, September-October.

NZS3101, 1995, Concrete Structures Standard, NZS3101, Parts 1 & 2, Standards New

Zealand, Wellington, New Zealand.

NZS3101, 2004, Amendment No. 3 to 1995 Standard (NZS3101), Standards New Zealand,

Wellington, New Zealand.

Page 1-17
Norton J.A, King A.B, Bull D.K, Chapman H.E, McVerry G.H, Larkin T.J and Spring K.C,

1994, Northridge Earthquake Reconnaissance Report, Bulletin of the NZNSEE, Vol.

27, No. 4, December.

OGrady C.R, 2004, Letters to the Editor, Journal of the New Zealand Structural Engineering

Society (SESOC), Vol 17, No. 2, pp 8-9.

Oliver S.J, 1998, The Performance of Concrete Topped Precast Hollow-core Flooring

Systems Reinforced with and without Dramix Steel Fibres under Simulated Seismic

Loading, Masters Thesis, Department of Civil Engineering, University of Canterbury,

Christchurch, Canterbury.

Priestley M.J, Sritharan S, Conley J.R, Pampanin S, 1999, Preliminary results and

conclusions from the PRESSS five-storey precast concrete test building, PCI Journal,

Vol. 44, No. 6, November-December, pp 42-57.

Restrepo J.I, Park R and Buchanan A, 1993, Seismic behaviour of connections between

precast concrete elements, Research Report 93-3, Department of Civil Engineering,

University of Canterbury, Christchurch, New Zealand.

Zerbe H.E, and Durrani A.J, 1989, Seismic Response of Connections in Two-bay R/C Frame

Subassemblies, Journal of Structural Engineering, Vol. 115, No. 11, November, pp

2829-2844.

Zerbe H.E, and Durrani A.J, 1990, Seismic Response of Connections in Two-bay R/C Frame

Subassemblies with a floor slab, ACI Structural Journal, Vol. 87, July-August, pp 406-

415.

Page 1-18
2. Section 2 Sub-Assemblage Experimental Investigation

SECTION SUMMARY

The present research experimentally investigates, by means of a two-dimensional

sub-assemblages, two hollow-core-to-support beam seated connections under simulated

seismic loads. The support connections are reinforced and rigidly tie the floor into the

supporting beam, the first of which has been proposed in Amendment No. 3 to the New

Zealand Concrete Standard, (NZS 3101, 2004). A newly developed test set-up is described

which incorporates both the relative rotation between the supporting beam and hollow-core

floor, and tension in the floor, to represent the elongation of a parallel moment resisting

frame, as the sources of damage. The sub-assemblage specimens were tested up to equivalent

inter-storey drifts of 5.0% and visual and instrumental observations from the experiments are

outlined. A simplified analytical study is conducted to better understand the performance and

observed failure modes. Results show promising behaviour, and after testing in a concurrent

full-scale super-assemblage experiment, recommendations for adoption into the forthcoming

revision of the New Zealand Concrete Standard, NZS 3101, are made.

Page 2-1
2.1 INTRODUCTION

Recent earthquake events and research undertaken has raised serious concerns of the seismic

performance of hollow-core floor systems. Experiments undertaken by Matthews (2004) at

the University of Canterbury have showed that a hollow-core floor diaphragm supported by a

ductile moment resisting concrete frame structure performed inadequately under simulated

seismic loads. Further experimental research conducted by Lindsay (2004) and Lindsay et al

(2004) demonstrated that improved detailing significantly improved the performance of these

hollow-core floor systems. Amendments to the New Zealand Concrete Design Standard (NZS

3101, 2004) stipulate two new acceptable solutions. The two connections could be

characterised as either pinned or rigid. The former pinned connection detail was tested

by Lindsay (2004), but the latter remains untested, and is the subject of this research.

In this section, the seismic performance of two hollow-core floor unit-to-support beam

connections is experimentally investigated. The experiments were conducted in a two-

dimensional sub-assemblage component test format. The loading set-up was developed from

previous sub-assemblage tests (Bull and Matthews, 2004 and Liew, 2004) whereby tensile

elongation effects were imposed on the floor-to-beam connection. This was considered

necessary to simulate the effect of beam growth that takes place in plastic hinge zones as a

result of cyclic loading effects. The two specimens feature rigidly reinforced connection

details, one of which is the detail proposed in Amendment No. 3 to NZS 3101:1995 (2004)

and the other is a modified detail following similar principles.

This section firstly reviews research relevant to these experiments, both at sub-

assemblage and super-assemblage formats and the progression made from this research to the

current specifications detailed in the New Zealand Concrete Design Standard. The details of

the test specimens and experimental set-up are provided and discussed. The second part of

this section outlines the results from the experiments.

Page 2-2
2.2 PREVIOUS RESEARCH FINDINGS AND NEW ZEALAND GUIDELINES

There have been several research programs at a sub-assemblage level that have investigated

hollow-core support conditions. Meija-McMaster and Park (1994), Herlihy et al (1995),

Oliver (1998) and Herlihy and Park (2000) examined seat requirements of hollow-core units

due to beam elongation. In each set of experiments, various end connection details were

investigated under either a monotonic or a cyclic horizontal load, to simulate the loss of

seating (pull-off) effects that occur due to beam elongation. The experiments were primarily

undertaken to investigate whether the cast in-situ topping slab and starter bar reinforcement as

well as different reinforced core configurations are sufficient to stop the floor units from

collapsing when the hollow-core unit was pulled off the seat of the supporting beam.

Meija-McMaster and Park (1994) demonstrated that various reinforced connections

between the hollow-core floor units and supporting beams, designed to carry the weight of the

floor units, performed well when a loss of seating occurred. The use plain round bars also

showed that a ductile connection could be achieved due to an early loss of bond strength

enabling large plastic elongations to be sustained. A limitation of the use of the beam-to-floor

connections used within structures expected to experience large horizontal movements was

placed. Herlihy et al (1995) conducted several experiments looking at traditional beam-to-

floor connection details used in New Zealand construction. The testing showed that the

hairpin type of connection (reinforcement hooked at both ends connecting the floor and

supporting beam at the bottom of the hollow-core cells) performed adequately but concluded

that there was insufficient composite bond strength between the cast-in-place topping and the

hollow-core units to maintain a ductile connection when there was no reinforcement crossing

the topping to hollow-core interface. The authors raised concern over the performance of the

connections within structures under large horizontal movements and proposed an alternative

connection detail, the paperclip using plain round bars (two layers of reinforcement in the

Page 2-3
shape of a paperclip between the hollow cores and supporting beam) that could be used to

resist these large displacements. Further research by Herlihy and Park (2000) included a

monotonic downward loading to induce tension in the topping and to observe the amount of

additional bending moment resistance added by the continuity steel.

Oliver (1998) examined the paperclip connection in more detail. Similar

conclusions were found to suggest that deformed bars traditionally placed in topping slabs are

insufficient to resist the effects of beam elongation that occurs in ductile frames. Oliver

(1998) concluded that the paperclip tie reinforced connection is not suitable in situations

where beam elongation is critical, such as the corner regions of a ductile concrete moment

resisting frame structure. The use of steel reinforced fibre concrete (SFRC) was also

investigated, and it was shown that the increased tensile capacity of the concrete due to the

steel fibres meant that higher displacements could be sustained.

The aforementioned research has examined the performance of hollow-core flooring

under the tension induced in the floor system from beam elongation effects. The failure

modes witnessed from these tests were different to the damage observed in the aftermath of

the Northridge Earthquake, when a loss of seating failure of hollow-core occurred (Norton et

al, 1994) and photographs showing the different failure modes are shown in Figure 2-1. The

relative rotation between the hollow-core floor slab units and the supporting beams was not

examined and therefore, failure due only to the induced tension is unrealistic.

Matthews (2004) tested in a full-scale three-dimensional super-assemblage, the

supporting beam-to-hollowcore slab connection details commonly used in New Zealand

construction prior to 2001. The connection detail consisted of hollow-core units seated on a

mortar bed on a 20mm or 40mm concrete seating. This mortar bed did not allow the floor

units to slide horizontally relative to the supporting beam as originally assumed. The cyclic

relative rotations between the floor and supporting beams, due to lateral loading of the super-

Page 2-4
assemblage, resulted in diagonal compression stresses from the seating to the topping slab and

orthogonal tensile stresses within the unreinforced webs of the hollow-core units. Due to the

cyclic loading, these tensile stresses led to cracking and ultimately, the loss of seating failure

of the hollow-core units (Figure 2-1(c)). The experimental results showed that the relative

rotation between the flooring and supporting beam was the major contributing factor to

failure.

(a) Meadows Apartment Building, Northridge (b) Herlihy and Park (2000) failure mode.
Earthquake 1994 (Norton et al, 1994).

(c) Matthews (2004) failure mode. (d) Liew (2004) Specimen 2 failure mode
Figure 2-1 Comparison of the failure modes observed during previous experiments.

One of the actions taken following the inadequate behaviour of the hollow-core floor system

observed during the Matthews (2004) experiment was the formation of a Technical Advisory

Group on precast floors (TAG). TAG sought to address the identified problems associated

with hollow-core floor systems and proposed new hollow-core support connection solutions

for future construction subject to experimental investigations.

Page 2-5
Bull and Matthews (2003) conducted four proof-of-concept sub-assemblage tests that were

commissioned by Precast NZ. The testing regime introduced the relative rotation between the

floor slab and support beam as the chief failure mode by way of a cyclic vertical load at the

end of the 6m long hollow-core units. A diagram of the test set-up is shown in Figure 2-2(d)

and the specific details for the test specimens are included in Figure 2-2.

(a) Control Specimen (200 and 300 series) (b) Flexible seating detail

(c) Paperclip seating connection detail (d) Test configuration

Figure 2-2 Seating connection details and the test configuration used by
Bull and Matthews (2003)

Two of the specimens comprised of traditional connection details, one with a 300 series

hollow-core unit and the other with a 200 series hollow-core unit, denoted as the control

specimens. The traditional connection details (for the 200 and 300 series hollow-core units)

comprised of the floor unit seated on a mortar bed, core end plugs to prevent concrete from

entering the cores, and conventional continuity starter bar reinforcement in the topping slab.

The floor-to-beam connection for the 300 series hollow-core unit was identical to the detail

used in the Matthews full-scale super-assemblage experiment (Matthews, 2004). The other

two specimens were newer connection details, recommended by TAG, and one of which

effectively created a pinned connection with a compressible backing board and a low

Page 2-6
friction bearing strip, and the other was a rigid paperclip connection, tying the beam and

floor slab together with reinforced cores. The traditional methods behaved poorly with web

cracking observed at low inter-storey drifts and full delamination of the hollow-core units

from the topping slab. There was also significant displacement of the floor units down the

face of the supporting beams at higher drifts. The damage patterns were similar between the

200 series and 300 series hollow-core specimens as well as what was observed in the full-

scale three-dimensional Matthews (2004) tests (Figure 2-1(c)). The two newer methods

behaved well and were able to sustain high drifts without a loss of seating occurring.

As a result of the previous research and the widespread use of precast floors in

construction practice, specific details governing the use of precast concrete systems have been

implemented within New Zealand design standards. Amendment No. 3 to NZS 3101:1995

(2004) provides two acceptable solutions for connecting hollow-core floors into supporting

beams, as indicated by Figure 2-3(a) and (b). The first solution is a pinned, simple type of

connection, which features the hollow-core unit seated on 75mm seat with a low friction-

bearing strip, and has a compressible backing board between the end face of the hollow-core

unit and supporting beam. This detail was experimentally tested and validated, firstly by Bull

and Matthews (2004) and then in a full-scale super-assemblage experiment (Lindsay, 2004).

Significantly improved behaviour was observed (as compared to Matthews (2004)), and the

beam-floor connection was able to withstand super-assemblage inter-storey drifts to +5%. The

second of these acceptable solutions has currently not been tested experimentally. This

floor-to-beam connection solution, as shown in Figure 2.3(b) could be classified as a rigid

type of connection with reinforcing tying the hollow-core units into the supporting beam. The

rigid reinforced detail requires more effort to construct (as compared with the simple

connection detail tested by Lindsay (2004)) with the need to pre-cut cores and place extra

reinforcement.

Page 2-7
(a) Hollow-core with compressible backing on low friction bearing strips.

(b) Hollow-core with 2 2 leg R 16 hairpins on low friction bearing strips.

(c) Capacity design actions in hollow-core.


Figure 2-3 Support for hollow-core flooring units: details proposed in Amendment
No. 3, NZS 3101:1995 (2004) (Figure C4.3/2).

Page 2-8
However, the rigid connection offers redundancy by being tied into the supporting beams, and

because of the chance for hollow-core units to arrive on site short (due to shrinkage, creep or

construction tolerances), this type of connection construction can be rendered more suitable.

Liew (2004) continued the research from Bull and Matthews (2003) and investigated

the relative rotation characteristics of hollow-core seated connections. Two new connection

details were investigated as well as a retrofit solution, which are all shown in Figure 2-4. The

first detail included a 70mm seating, and each core (4 cores) of the hollow-core unit was

reinforced with Grade 500 R16 reinforcement and Grade 500 D12 starter bars (where the

R and D denote plain round and deformed reinforcement respectively). The second

specimen was identical to the first specimen, except that there was no seating. This negative

seating case was to investigate whether units that arrive short to site could be propped and

connected to the supporting beam with a strongly reinforced bridging connection. The third

specimen used the same traditional detail previously tested by Matthews (2004) and Bull and

Matthews (2003) but featured an angle section catcher to act as a retrofit measure.

The findings from Liews (2004) experiments highlighted that over-reinforced beam-

to-floor connections leads to inferior performance. Incipient failure of Specimen 1 occurred at

a drift of 1%, and despite suffering significant damage, Specimen 1 was able to function

without collapse up to 4% drifts. The detail however performed poorly in comparison to the

paperclip detail tested under the Bull and Matthews (2004) research program (Figure 2-2(c)),

which utilised two D12 (Grade 300) paperclips instead of the four R16 (Grade 500)

paperclips. The zero seating specimen performed inadequately and failed under the expected

2% drift design demand. Although the moment capacity of the connection was suitable, the

lack of shear resistance from the zero bearing proved to be totally inadequate. The retrofit

detail tested did not perform as well as the author would have hoped, with the angle bracket

effectively catching the unit, but at the same time restraining the rotation of the floor unit.

Page 2-9
Hurricane "Ductile Mesh"
HD12 starter bars
612 100
D24

43
R60 Plastic end plug

4 R16 Paperclips

65
H/C unit seated on mortar pad

70

(a) Specimen 1: Paperclip detail with 70mm seating


HD12 starter bars Hurricane "Ductile
612 100
Mesh"
43

R60 Plastic end plug

4 R16 Paperclips
65

(b) Specimen 2: Paperclip detail with zero seating


R 1 2 @ 1 0m m c/c H R C 66 5 M esh
612

P la stic e nd p lug

1 70 150 x150 x12 E .A


16

M cD ow el B ea ring S trip seate d on A ng le S eat


A nch or B olt
20

(c) Specimen 3: Retrofit of a grouted seat


Figure 2-4 Test specimens and their dimensions from testing by Liew (2004).

Liew (2004) concluded that it would be interesting to repeat the experiments including the

effects of beam elongation. In the early sub-assemblage research performed, only horizontal

pull-off loads were used. The sub-assemblage experiments by Bull and Matthews (2003) and

Liew (2004) included cyclic vertical loading inducing relative rotations between the hollow-

core floor unit and the supporting beams. To this point, no sub-assemblage experimental

Page 2-10
program has investigated the effect of both horizontal tension in the floor and relative rotation

between the floor unit and supporting beam.

The second hollow-core connection detail stipulated by Amendment No. 3 to NZS

3101:1995 (2004), as shown in Figure 2-3(b), currently has not been experimentally tested.

For this reason, this connection detail between the supporting beam and hollow-core is the

focus of this experimental study.

2.3 TEST SPECIMEN DETAILS

The two test specimens are described in the following subsection and can be seen in Figure

2-5(a) and (b). Full design and construction details of the specimens are provided in

Appendix A.

2.3.1 Specimen 1

The seated connection detail used for this experiment is the second detail proposed by

Amendment No. 3 to NZS 3101:1995. The 6m long 300 series hollow-core unit was seated on

a 75mm wide seat and a low friction bearing strip. The connection features two of the four

hollow cores reinforced and filled with concrete. Starter bars consisted of D12 (Grade 300)

reinforcement spaced at 300mm centres. Two cores of the four hollow cores were reinforced

with hooked R16 (Grade 300) bars placed close to the bottom of the cores. The topping slab

consisted of D12 (Grade 300) reinforcement, which was lapped with the starter bars, and

placed at 300mm centres in both directions. It is also worth noting that backing boards were

used in place of traditionally used end plugs, to prevent concrete entering the un-filled cores.

This is to aid the beam-to-slab interface becoming the critical zone, concentrating the

rotational demand at this interface and not into the hollow-core unit.

Page 2-11
D12 Starter Bars (Grade 300)

2*R16 (Grade 300) - 1 per filled core


Seat Transverse Low Friction
Reinforcement bearing strip
75mm seat

(a) Code Amendment detail

D12 Starter Bars (Grade 500) @ 300mm,


continuous as topping reinforcing

4*D12 (Grade 500) - 2 per filled core


Seat Transverse Low Friction
Reinforcement bearing strip
75mm seat

(b) Modified seating detail


Figure 2-5 Test specimen details.

2.3.2 Specimen 2

Specimen 2 was based principally on Specimen 1. However the Grade 300E reinforcement

was replaced with Grade 500E deformed reinforcement to represent present construction

practice in New Zealand. The hollow-core unit is similarly seated on a 75mm seat and a low

friction bearing strip, and features two reinforced cores. Starter bar reinforcement consisted of

D12 bars placed at 300mm centres above each of the cores. The R16 reinforcement used for

Specimen 1 in two of the four cores has been replaced by 2-D12 bars for Specimen 2. The

reason for this was to provide equal top and bottom reinforcement (4-D12) and hence equal

Page 2-12
positive and negative connection moment capacities. Transverse reinforcement was used

within the reinforced cores, with the placement of one leg of R10 reinforcement at 150mm

spacings, tied between the starter bar and bottom reinforcement. This was to investigate

whether implementation of transverse reinforcement in hollow-core units is an adequate way

to overcome situations when overly high shear demands are present, which might be

experienced with long spans, bridges or other cases. The reinforcement passing the beam-to-

slab interface (4-D12 reinforcement top and bottom) has been de-bonded over 150mm with

plastic tubing to reduce the risk of reinforcement fracturing and to support the critical section

occurring at the beam-slab interface. In addition to this, a thin plastic strip was placed on the

top of the starter bars beneath the concrete surface to both act as a crack initiator and further

promote rotation at the critical section.

2.4 BEAM ELONGATION

The inelastic rotations experienced by plastic hinges of concrete frame structures under

seismic loading cause the phenomenon of beam elongation, whereby the beam grows in

length. Beam elongation has been quantified through many research programs over the last

decade such as Fenwick et al (1999). Typical values for beam elongation vary from between

2% and 5% of the beam depth, and these estimated values have been used to explain

permanent deformations of frame structures as well as a way of assessing precast floor seating

widths.

Matthews (2004) and Matthews et al (2004) presented a new theory for predicting

elongation for structural concrete members. The theorem is based on simple rigid body

kinematics and utilises a rainflow counting method to calculate the amount of elongation due

to inelastic bending of a plastic hinge in a frame system. Elongation occurs only if plastic

Page 2-13
deformations exceed those of previous peaks. In its essence, the theorem can be given by the

expression:

Gi T u ecr (2.1)

in which Gi = elongation of the ith hinge; T = rotation of the hinge; and ecr = eccentricity

between the centre of gravity of the concrete (c.g.c) for the beam and the centroid of the

compression force (instantaneous centre of rotation, I.C.R.).

The elongation of a frame (Gel) with ith plastic hinges can be separated into the elastic

and inelastic components and can be expressed as:

T e
n
G el G eel  G pel 
p  T p  T y cr (2.2)
i 1

where G eel = unloading recoverable elastic elongation, G pel = non-recoverable plastic

elongation, T p = positive plastic rotation, T p = negative plastic rotation, T y = yield rotation of

the frame.

The growth of the beam lengths occurs as tensile strains within the reinforcement are

higher than the accompanying compression strains within the concrete. Under reversing

loading, compression yielding does not fully close the flexural cracks caused by the

reinforcement which has previously yielded in tension. Figure 2-6(a) illustrates beam

elongation by showing the relationship between the hysteretic behaviour and recoverable

elastic and non-recoverable plastic components of elongation. The theorem is dependent on

the loading, with respect to the plastic rotations and the yield drift and ecr values are a

function of the structural and geometrical properties of the structure. The ecr eccentricity value

for example is a function of the reinforcement and the presence of topping reinforcement and

prestressing strands of adjoining floor slabs. Figure 2-6(b) and (c) show the derivation of the

ecr values for a typical concrete frame. For a more in-depth explanation of the development of

this model, reference can be made to Matthews (2004).

Page 2-14
2.5.1 Sub-Assemblage Elongation

With the beam elongation prediction model proposed by Matthews et al (2004) and further

verified by Lindsay (2004), there became a method to introduce beam elongation effects into

two-dimensional sub-assemblage tests. A second actuator has been implemented to induce

tension forces in the floor through a displacement controlled loading, which was modelled on

the beam elongation prediction model. The drift displacement history was controlled by the

vertical actuator in a prescribed manner, and as the theorem predicts beam elongation in terms

of rotation (of a plastic hinge or frame) under lateral loading, the use of Equation 2.2 allowed

beam elongation results to be calculated for each inter-storey drift increment. After

accounting for geometry (as the horizontal actuator rotates), an input file was developed

listing the target ram displacements for both the vertical and horizontal rams at each

increment of drift.

The amount of elongation was calculated with reference to the super-assemblage

frame used during the Matthews (2004) project as well as the full-scale tests associated with

this current research. Beam elongation was calculated as one half of the total frame

elongation, or the pull-off displacements which one end support of the hollow-core flooring

units would experience. The amount of beam elongation calculated was based on an exterior

moment resisting frame and it is accepted that there would be a difference between this value

and that experienced by a floor unit that is seated away from the frame due to out-of-plane

bending and torsion of the transverse supporting beams. Also, a yield drift rotation, T y , of

0.5% was assumed and used for the generation of the beam elongation data for these

experiments. It is later verified that although this was a priori assumption, it was a valid

approximation.

Page 2-15
Load
3 5

Deflection

'el T+p

Unloading
Tp- 5
recovery
4
Unloading
recovery
3
Elastic Elastic
growth growth
2 1

T y Ty DriftT

(a) Detailed diagram of beam elongation showing the hysteretic relationship and recoverable
elastic and non-recoverable plastic components.
T

I.C.R Elongation due to


negative moment

ecr |0.475D D c.g.c


jD jD
ecr |0.425D

Elongation due to
positive moment

Instantaneous centre
of rotation (I.C.R)

(b) Exterior plastic hinge lever arms for reinforced concrete beams
I.C.R
Elongation due to T
negative moment
Prestressing
strands

c.g.s
ecr |0.275D D
jD
ecr |0.225D c.g.c

Elongation due to
positive moment

Instantaneous centre
of rotation (I.C.R)

(c) Interior plastic hinge lever arms

Figure 2-6 Diagrammatic representation of beam elongation (adapted from Matthews


(2004))

Page 2-16
2.5 EXPERIMENTAL SET-UP

The construction of the test specimens, albeit sub-assembly components of a real structure,

was conducted in accordance with customary site practice. Figures 2.7(a) and (b) shows a few

stages of the construction process. Initially, the support beams were cast to half height using

typical reinforcement for a concrete frame beams. The 6m long 300 series hollow-core units

were then lowered into place on top of a low friction bearing strip. The top flanges of two of

the hollow-core voids were pre-cut 75mm wide and approximately 800mm long to allow for

the placement of the connection reinforcement. The connection and topping reinforcement

was put in place together with cast in beams to tie down the support beam, and the top half

beam, reinforced cores and 75mm topping concrete were cast.

Tensile tests were carried out on samples of the reinforcement used in the two tests.

For Specimen 1, the average yield strengths were 311 MPa for the D12 reinforcement and 325

MPa for the R16 reinforcement. For Specimen 2, the average yield strength for the D12

(Grade 500) reinforcement was 570 MPa. Standard structural ready mix concrete was used

with a specified 30 MPa 28-day strength, 19mm maximum aggregate size and a 100mm

slump. Test day compression cylinder test strengths were 41 MPa and 35 MPa for the topping

concrete of Specimens 1 and 2 respectively, and 45 MPa for the supporting beams. Further

information from the material tests is provided in Appendix A.

The configuration of instruments used in the tests is illustrated in Figure 2-7(c). Strain

gauges were attached to the starter bar and hairpin bar reinforcement to measure the level

and amount of propagation of inelastic behaviour into the topping slab and hollow-core unit.

Linear potentiometers were used to measure the relative movement of the supporting beam

and floor slab as well as the horizontal pull-off and vertical dropping of the hollow-core unit

relative to the supporting beam. Inclinometers were used to attain an accurate level of the drift

Page 2-17
imposed on the floor. Purpose-built data acquisition and control programs were employed to

track the experimental data and automatically control the two actuators loading the specimen.

A new loading rig and set-up was developed to test the two specimens. This loading

set-up, through the use of dual displacement controlled actuators, loads the specimen in such a

way that both the relative rotation between the floor and supporting beam and induced tension

in the floor are integrated. A schematic of the set-up is shown in Figure 2-7(d). In the same

way as the Bull and Matthews (2003), and Liew (2004) sub-assemblage tests were

undertaken, a vertical actuator was located at the end of the 6m long hollow-core unit. This is

able to provide negative and positive moments at the connection of the supporting beam with

the hollow-core unit. In addition, a 35kN concrete block is placed at mid-span to generate

extra shear force at the connection, which assures an equivalent shear force distribution to the

full-scale prototype structure. The development from the previously conducted sub-

assemblage tests (outlined in Section 2.2) is the coupling of the motions of the vertical and

horizontal actuators. The horizontal actuator induces a net tension force in the floor slab and

mimics the effects of beam elongation that occurs in three-dimensional structures.

The displacement controlled loading history of the vertical actuator comprised of 2

complete reversing cycles to inter-storey drift amplitudes of 0.5%, 1%, 2%, 3%, 4%,

and 5%. The inter-storey drift used throughout this section refers to the rotation (as a

percentage) of the hollow-core floor unit with respect to the supporting beam. Positive drift

denotes tension on the bottom of the hollow-core unit and negative drifts imply tensions

within the starter bars and topping concrete of the floor unit. It should also be noted that from

analysis of the test data, the target drifts and displacements were not always achieved due to

the rotation and sliding of the supporting beam relative to the ground. Throughout the results

and analysis presented in the following sections, this discrepancy has been accounted for.

Page 2-18
(a) Photograph showing the pre-cutting of the (b) Supporting beam reinforcing and
hollow core. connection detail (Specimen 1).

Starter bar strain gauges

Inclinometers (spaced
along the floor unit)

Beam-to-floor Hairpin bar strain gauges


potentiometers

Floor unit pull-off


potentiomenters
Floor unit vertical
displacement
potentiomenters

(c) Instrumentation layout

Vertical
(drift controlling) Reaction Frame
Actuator

Reaction Frame

35kN Concrete
Weight

Support beam and


restraints

300 series hollow-core unit

6000mm Horizontal Hollow-core unit


(beam elongation) Support beam and
Actuator restraints

South (side) elevation West (end) elevation

(d) Experiment set-up: location of the actuators and reaction frame.

Figure 2-7 Test set-up photographs and diagrams

Page 2-19
2.6 EXPERIMENTAL RESULTS: SPECIMEN 1

Figure 2-8 presents photographs of damage observed for the testing of Specimen 1 and further

results are displayed in Appendix A. A crack opened at the floor unit-to-supporting beam

interface during the 0.5% cycle and this widened to around 5-6mm during the 1% cycle

peaks. During the 2% cycles, this continuity crack had opened to around 10mm on the

tension side and 5-6mm on the compression side, and following the second cycle to 2%, the

residual opening at this interface was 9-10mm, which signified the amount of beam

elongation pulling the floor unit off the seat of the supporting beam. To this point, the only

other signs of damage were hairline cracks that opened up at 2% within the topping concrete.

However, on the subsequent cycle these cracks closed and did not reopen throughout the

remaining testing. The 3% and 4% cycles were characterised by the continuing widening of

the beam-floor opening, and at the conclusion of the test after a second 4% cycle, the

residual opening was around 25mm.

The performance of this specimen is easier to describe by discussing what did not

occur during the test. The large crack opening at the beam-floor interface showed that the

rotational demand imposed on the floor unit was nil, and no damage of the floor unit was

observed. The bearing strip performed its intended purpose of allowing the unit to slide

horizontally relative to the supporting beam while staying in place on the concrete seat. The

improved performance of the bearing strip also meant that no spalling of the concrete seat of

the supporting beam occurred.

Page 2-20
(a) Cracking of topping at 2%. (b) Beam-to-floor interface crack at the end of
testing.
Figure 2-8 Photographs of the testing of Specimen 1.

A collection of the important instrumental results is shown in Figure 2-9 for Specimen 1. The

overall hysteretic performance of the specimen is shown by the moment rotation relationship

in Figure 2-9(a). The hysteresis shows good energy dissipation of the beam-to-floor

connection at the interface and no premature degradation of stiffness or strength to drifts of

4%. The maximum positive and negative moment capacities occurred at the first cycle to

0.5% respectively, and the following cycles to 4% showed a constant level of moment

strength. It is noticeable for the negative drift cycles, that there were reductions in strength

between the first and second cycle. The theoretical predictions show good agreement, and the

derivation of the nominal and overstrength theoretical values are outlined in Section 2.8. The

overstrength was determined from material testing for the reinforcement, and was taken to be

the ratio of the ultimate tensile stress to the yield tensile stress. Behaviour from the strain

gauged reinforcing is presented in Figures 2-9(b) through (e). The notation of the north core

and south core refers to the northern most and southern most reinforced cores, or the second

and fourth cores. Each of the four graphs for the starter and hairpin show similar trends.

Page 2-21
100

Moment
(kNm)
80

60

40

20

0
-5 -4 -3 -2 -1 0 1 2 3 4 Drift (%) 5
-20

-40

-60 Experiment
Nominal Capacity
-80 125% Overstrength Capacity

-100

(a) Moment rotation hysteresis performance of Specimen 1.


Starter bar

1 2 3
1 & 4: The drift displacement peak (time)
2 & 5: denotes is a count of the half cycles.
4 5 6
3 & 6: Appendix A provides a list of the
Yield: displacement peaks for each count
Hairpin bar

45 45

40 40

35 35

30 30
strain
strain

25 25

20 20

15 15

10 10

5 5

0 0
0 10 20 30 40 0 10 20 30 40
-5 -5
Drift displacem ent peak (tim e) Drift displacem ent peak (tim e)
(b) North core starter bar strains. (c) South core starter bar strains.
50 45
45 40
40 35
35
30
30
strain

strain

25
25
20
20
15 15

10 10

5 5
0 0
-5 0 10 20 30 40 0 10 20 30 40
Drift displacem ent peak (tim e) -5
Drift displacem ent peak (tim e)

(d) North core hairpin bar strains. (e) South core hairpin bar strains.

Figure 2-9 Experimental Results for specimen 1: code amendment detail.

Page 2-22
The strain measured by the strain gauges at the critical section at the beam-floor interface

showed distinctly higher strains. This matches the experimental observations whereby the

damage and cracking was concentrated at the beam-to-floor interface, and there was minimal

amounts of cracking into the topping slab over the hollow-core unit. This indicates that the

design objective of concentrating the rotational demand at the critical section was achieved.

For the starter bar reinforcement, yielding occurred early either at 0.5% or during the first

cycle to -2% at the floor to beam interface. Whereas yield penetration reached the topping

strain gauges 150mm away during the subsequent 2% cycles. Yield penetration however did

not progress back into the beam until the final stages of the test. The same trend is shown for

the hairpin reinforcement, where the plastic strains were highest at the critical section.

2.7 EXPERIMENTAL RESULTS: SPECIMEN 2

A collection of photographs from the testing of Specimen 2 are presented in Figure 2-10 and

extended results are shown in Appendix A. Very little damage occurred during the 0.5%

cycles where there was only cracking at the sides at the end of the floor unit of the beam-slab

interface and minor cracking within the topping slab. At 1% the continuity crack along the

interface between the floor slab and supporting beam had opened to around 4mm, at +2% it

had closed to be a hairline crack and at 2% it was 8-10mm wide. During the second cycle to

+2%, spalling of the seat concrete of the supporting beam was observed which progressively

worsened throughout the remainder of testing. At the second cycle to 3%, the continuity

crack had widened to be 20mm at both the top and bottom of the hollow-core unit (Figure

2-10(b)). On the first cycle to +4%, a loud noise and accompanying drop off in load-carrying

capacity was noticed at +3.8% suggesting reinforcement fracture of one of the bottom D12

bars. On the subsequent second cycle to 4% there were similar fractures at +2.84% and then

at 0.9%, -2.2% and -3.3%. Following the 4% cycles, the spalling had extended over the

Page 2-23
whole width of bearing of the floor unit. The bearing strip had become clearly visible as can

be seen in Figure 2-10(c), and the seat width had been reduced to approximately 20mm. There

was also evidence of the floor unit dropping down the face of the supporting beam. Prior to

the 4% cycles the damage was moderate and repairable. However, following the first cycle

to +4% the damage could be categorised as major and a similarly damaged building could

only be entered under extreme caution. At +5% the bearing strip had split, was sliding

horizontally off the seat with the hollow-core unit, and the unit had dropped by around 10mm

down the face of the supporting beam. The beam-floor crack had opened to widths over

35mm. At +4.8% on the second cycle to +5% the north starter bar fractured and the seated

support was lost.

(a) Topping and beam-floor interface cracking (b) Beam-floor interface crack at
at 2%. +3% drift.

(c) Extent of seat concrete spalling. (d) Failure after a loss of seating (at 4.78% on
the second cycle to 5%.
Figure 2-10 Photographs of the testing of Specimen 2

Page 2-24
Figure 2-11 displays the key results derived from the instrumentation from the testing of

Specimen 2. The moment rotation behaviour is illustrated in Figure 2.11(a). An additional

20kN concrete weight was added at the midspan for upward loading (unloading from a

negative drift peak), and both midspan weights taken off for downward loading (unloading

from a positive drift peak to a negative drift peak). The increased bending moment gradient

heightened the shear force at the support imparting higher shear demands on the beam-floor

connection. In this case the compatibility with the full-scale testing was forsaken in order to

more rigorously test the shear reinforcement of the connection detail.

The instances of the reinforcement fracture are shown on Figure 2-10(a) by the sudden

decrease of load carrying capacity. The overall hysteresis shows good performance up to the

first cycle to +4%, but after the first reinforcement fracture (at 3.83% on the first cycle to

+4%), there was a significant reduction in stiffness and load carrying capacity of the floor-to-

beam connection. There was a distinct reduction of peak moments between the first and

second cycles to +4%, largely due to reinforcement fracture at +2.8%, and on the second

cycle to 4%, a series of three reinforcing fractures reduced the load carrying capacity

appreciably. The capacity of the floor-to-beam connection during the 5% cycles was

minimal as a result of the prior reinforcement fractures, and overall failure occurred when the

final starter bar fractured, under a positive drift, and the floor unit lost its seating at +4.78%

on the second to +5%.

The maximum positive moment capacity of the beam-floor connection occurred at

+4%, and the maximum negative moment capacity occurred at 0.5%. This shows that for the

positive drift cycles there was a good degree of strain hardening of the bottom reinforcement,

which would have occurred prior to the reinforcement fracture. The theoretical predictions

overlain on the moment rotation relationship show good agreement with the experimental

results.

Page 2-25
100

Moment
(kNm)
80
60
First
40
Fracture
20
Drift (%)
0
-6 -5 -4 -3 -2 -1 -20 0 1 2 3 4 5 6

-40
Fracture
Fracture
-60 Experiment
Nominal Capacity
-80 125% Overstrength Capacity
-100

(a) Moment rotation hysteresis performance of specimen 2.


Starter bar

1 2 3
1 & 4: The drift displacement peak (time)
6
2 & 5: denotes is a count of the half cycles.
4 5
3 & 6:
Appendix A provides a list of the
Yield:
displacement peaks for each count
Hairpin bar

45 45
40 40
35 35
30
30
25
strain

25
strain

20
20
15
15
10
10
5
0 5

-5 0 10 20 30 40 0
0 10 20 30 40
-10 -5
Drift displacem ent peak (tim e) Drift displacem ent peak (tim e)

(b) North core starter bar strains. (c) South core starter bar strains.
40 45

35 40

30 35

30
25
25
strain

strain

20
20
15
15
10
10
5
5
0 0
0 10 20 30 40 0 10 20 30 40
-5 -5
Drift displacem ent peak (tim e) Drift displacem ent peak (tim e)

(d) North core hairpin bar strains. (e) South core hairpin bar strains.

Figure 2-11 Experimental Results for specimen 2: modified detail.

Page 2-26
A comparison between the moment rotation behaviour for Specimen 1 and Specimen 2 shows

the relative strengths of the two reinforced connection details. For Specimen 2, the higher

positive moment capacity of the floor-to-beam connection and extra mid-span loads under

upward loading would have imparted larger compression reaction forces on the seat, and the

higher stresses can help to explain the amount of spalling.

Figures 2-11(b) and (c) present the strain behaviour of the starter bars, and similarly

2-11(d) and (e) the hairpin reinforcement for the floor-to-beam connection. The behaviour is

similar to that observed for Specimen 1, where the starter and hairpin reinforcement show

similar trends. Pre-yield compression strains were experienced during the early stages of the

test, however, following tensile yielding, the strains never returned below this yield strain

value and the reinforcement hysteretic behaviour would have been characterised by a distinct

shift into positive tensile strains. Yielding occurred early for the starter bars during the first

cycle to -1%, whereas yield penetration reached the topping concrete strain gauges (150mm

from the end of the beam-floor interface) during the 2% cycles and 150mm back into the

beam later during the 3% cycles. Plastic strains either side of the critical section of the

hairpin reinforcement were experienced later relative to Specimen 1.

To investigate the effects of vertical acceleration during seismic excitation, the testing

procedure could have been modified. Vertical accelerations have important implications, and

are accommodated for in design standards. Upward accelerations reduce the gravity loads on

the flooring, causing a reduction of the bending moment gradient, and extending the demand

on the topping reinforcement. The use of the additional concrete weight could have been used

oppositely (extra midspan loads for downward loading, and no loads for upward loading).

However, the strain relationship exhibited by the starter bar reinforcement shows that

maximum demand is placed on the reinforcement at the interface, and the strain reduces

further into the topping. The extent of which the strain diminishes away from the supporting

Page 2-27
beam over a greater length was not fully investigated, but it is assumed that either the

continuous starter bars, or starter bars lapped with the same type of diaphragm reinforcement,

that these effects would be accommodated.

2.8 STRENGTH AND FAILURE MODE PREDICTION MODELS

2.8.1 Moment capacity

To assess the strength of the reinforced floor-to supporting beam connections for both

specimens, simplified moment equations have been used in order to calculate the positive and

negative moment capacities. The structural actions imposed on the connection are shown in

Figure 2-3(c) and the moment capacities were calculated from the common reinforced

concrete relationship:

Mn = Asfy(d-d) (2.3)

in which As = the cross sectional area of tension reinforcement, fy = yield stress of tension

reinforcement, d = effective depth to the centre of the tensile reinforcement taken from the

outermost compression fibre, and d = depth from the outermost compression fibre to the

centre of the compression reinforcement.

Actual values for the yield stress and reinforcment and the measured lever arm

between the top and bottom reinforcement were used in the calculations. The overstrength

moment capacity of the two reinforced floor-to-beam connections was also calculated as

125% of the nominal (yield strength measured from material tests) strength. Table 2-1

presents the values used and the corresponding moment capacities:

The moment capacities have been plotted overlaying the experimental moment

rotation behaviour and can be seen in Figures 2-9(a) and 2-11(a). It can be seen that there is

good agreement with the experimental data for both of the test specimens. The moment

rotation relationship for Specimen 1 is illustrated in Figure 2-9(a) where it can be seen that

Page 2-28
the maximum positive and negative moments were experienced during the first cycle to

0.5%. Despite this, the predictions for Specimen 1 showed that the experimental values lay

within the 125% overstrength and nominal strength bounds. A similar quality of prediction is

evident for Specimen 2, as shown in Figure 2-11(a). Although the first cycle to 0.5%

produced the highest negative moment, there was some degree of overstrength or strain

hardening that was predicted by the calculated values. From these results, it is evident that the

use of an additional 25% strength to represent strain hardening effects is a satisfactory

indicator of component overstrength at the floor unit-to-seat connection.

Table 2-1 Nominal moment capacity calculations.

property f y (D12) f y (R16) As+ As- (d-d') Io M n+ M n-


units MPa MPa mm 2 mm 2 mm kNm kNm
Specimen 1 321 333 402 452 280 1.25 38 41
Specimen 2 570 - 452 452 280 1.25 72 72

2.8.2 Low cycle fatigue

When assessing the ultimate deformation capacity of a structure, low cycle fatigue of

reinforcing bars must be considered a likely failure mode, as described by Mander et al

(1994). Further investigation by Dutta and Mander (2001) produced a low cycle fatigue

theory that enables the plastic rotation life capacity of reinforcement to be calculated. The

theory is expressed as follows:

D L p 1 L p
Tp 0.16 1  2 N f
 0.5
(2.4)
D' D 2 L

in which Tp = plastic drift; L = lever arm of the cantilever column; D = overall member depth;

D = distance between outer layers of steel, 2Nf = the number of reversals to the appearance of

the first fatigue crack with which Nf is the effective number of constant amplitude cycles for a

Page 2-29
variable amplitude displacement history and Lp = equivalent plastic hinge length, or the length

of reinforcement where plastic strains are expected.

Miners well-known cycle counting method can be utilised to calculate the equivalent

number of equi-amplitude cycles (Nf):

2
T
Nf i T

(2.5)
ref

where T = drift for the ith cycle of loading and Tref = reference drift for an equivalent constant

amplitude loading history. A more complete description of this theory and the calculations are

present in Appendix A.

Using a reference drift of 4% an applied cyclic demand of Nf = 3.8 cycles is

calculated. Using Equation 2.4 and an equivalent plastic hinge length of the un-bonded length

of reinforcment (150mm), a plastic drift Tp of 0.03 radians was found. Together with an

elastic yield rotation of 0.7% that was derived from the experimental observations, the

rotational capacity of the reinforcement before fracture is 0.037 radians, or 3.7% as calculated

by Equation 2.4. This compares well to the observed first fracture that occurred at +3.83%

drift.

2.9 DISCUSSION OF RESULTS

The visual and instrumental observations show that both specimens behaved well. Specimen 1

was able to sustain drifts of 4% with no major damage to the supporting beam and hollow-

core floor slab. Specimen 2 performed well up to the low cycle fatigue fracture of the

reinforcement at +3.83% drift, but was still able to withstand further 4% and 5% cycles

until a loss of seating failure occurred at +4.78% on the second cycle to +5%. The relative

rotation damage was concentrated at the critical section at the interface between the

Page 2-30
supporting beam and floor interface for both specimens, and there was very little damage to

the hollow-core units and topping concrete away from this critical section. However, there

were still some shortcomings observed during the testing of Specimen 2. Large amounts of

seat spalling was observed, which minimised the bearing area for the floor unit. During the

final stages of testing, the amount of deterioration of the concrete seat meant that the hollow-

core floor unit was sliding down the face of the supporting beam. The low friction bearing

strip performed well for Specimen 1, but inadequately for Specimen 2, whereby the bearing

strip slid horizontally with the hollow-core floor unit relative to the supporting beam. The un-

bonding of the reinforcement passing the critical section for Specimen 2 also showed an

inability to prevent failure of the reinforcement.

The maximum moments sustained by Specimen 2 were 70% greater than the

maximum moments experienced by Specimen 1, which is to be expected from similarly

reinforced connections with Grade 300 and 500E reinforcement. The larger moment capacity

would have caused larger vertical reactions to be resisted by the seating concrete of the

supporting beam. This can help to explain why the bearing strip performed as intended for

Specimen 1, but not for Specimen 2, where the bearing strip slid horizontally with the hollow-

core unit relative to the seat of the supporting beam. In turn, the poor performance of the

bearing strip would have accelerated the spalling of supporting beam concrete seat. It is

difficult to infer at what point a floor-to-beam connection becomes over-reinforced, but the

relative damage between the two specimens indicates that the Grade 500E reinforced

connection may have been too strong. Also, the plain round R16 bars (used for Specimen 1)

appear to be more suitable in high ductility connections under high drift demands than the

deformed bars used for Specimen 2.

Page 2-31
2.10 CONCLUSIONS

From the experimental investigation presented in this chapter, the following conclusions may

be drawn:

1. The second of two hollow-core support details, as proposed by Amendment No. 3 to

NZS 3101:1995: a rigid, reinforced floor-to-beam connection, behaved well, and was

able to sustain inter-storey drifts up to 4% with little damage occurring to the

supporting beam and hollow-core unit.

2. A new testing procedure that introduced the realistic coupling of relative rotation of the

hollow-core and supporting beam and the net tension in the floor slab from beam

elongation three-dimensional effects as the main sources of damage was undertaken

successfully.

3. Simple analytical methods were proposed to assess the strength capacity of the hollow-

core to supporting beam connections. Similarly low cycle fatigue theory was used to

predict the rotation limit leading to reinforcement fracture. Both analyses provided

accurate representations when compared with the experimental data.

4. Evidence of over-reinforcement was observed with the use of Grade 500E reinforcement

with increased amounts of damage being observed for Specimen 2. The presence of plain

round reinforcement was more capable of resisting high ductility demands than the

unbonding of deformed reinforcement, which led to fracturing of the reinforcement.

2.11 REFERENCES

Bull D.K, Matthews J.G, 2003, Proof of Concept Tests for Hollow-core Floor Unit

Connections, Precast NZ report, Feb 2003.

Dutta A and Mander J.B, 2001, Energy Based Methodology for Ductile Design of Concrete

Columns, Journal of Structural Engineering, Vol. 127(12), December, pp 1374-1381.


Page 2-32
Fenwick R.C, Davidson B.J, and Lau D, 1999, Strength Enhancement of Beams in Ductile

Seismic Resistant Frames due to Prestressed Components in Floor Slabs, Journal of

the New Zealand Structural Engineering Society (SESOC), Vol 12, No. 1, pp 35-40.

Herlihy M.D, Park R and Bull D.K, 1995, Precast Concrete Floor Unit Support and

Continuity, Research Report 93/103, Department of Civil Engineering, University of

Canterbury, Christchurch, New Zealand.

Herlihy M.D and Park R 2000, Precast concrete floor support and diaphragm action,

Research Report 2000/13, Department of Civil Engineering, University of Canterbury,

Christchurch, New Zealand.

Lindsay R.A, 2004, Experiments on the seismic performance of hollow-core floor systems in

precast concrete building, Masters Thesis, Department of Civil Engineering,

University of Canterbury, Christchurch, New Zealand.

Lindsay R.A, Mander J.B, and Bull D.K, 2004a, Preliminary results from experiments on

hollow-core floor systems in precast concrete buildings, 2004 NZSEE Conference,

Rotorua, NZ, Paper 33, 8 pp.

Lindsay R.A, Mander J.B, and Bull D.K, 2004b, Experiments of the Seismic Performance of

Hollow-core Floor Systems in Precast concrete Buildings, 13th World conference on

Earthquake Engineering, Vancouver, Canada, CD ROM paper No. 585.

Liew H.Y, 2004, Performance of Hollow-core Floor Seating Connection Details, Masters

Thesis, Department of Civil Engineering, University of Canterbury, Christchurch,

New Zealand.

Mander J.B, Panthaki F.D, Kasalanati A, 1994, Low Cycle Fatigue Behaviour of Reinforcing

Steel, Journal of Materials in Civil Engineering, ASCE, Vol 6(4), November, pp 453-

468.

Page 2-33
Matthews J.G, 2004, Hollow-core floor slab performance following a severe earthquake, PhD

Thesis, Department of Civil Engineering, University of Canterbury, Christchurch,

New Zealand.

Matthews J.G, Mander J.B, Bull D.K, 2004, Prediction of beam elongation in structural

concrete members using a rainflow method, 2004 NZSEE Conference, Rotorua, NZ,

Paper 27, 8 pp.

Meija-McMaster J.C and Park R, 1994, Tests on Special Reinforcement for the End Support

of Hollow-core Slabs, PCI Journal, Vol 39, No. 5, pp. 90-105, September-October.

NZS3101, 1995, Concrete Structures Standard, NZS3101, Parts 1 & 2, Standards New

Zealand, Wellington, New Zealand.

NZS3101, 2004, Amendment No. 3 to 1995 Standard (NZS3101), Standards New Zealand,

Wellington, New Zealand.

Norton J.A, King A.B, Bull D.K, Chapman H.E, McVerry G.H, Larkin T.J and Spring K.C,

1994, Northridge Earthquake Reconnaissance Report, Bulletin of the NZNSEE, Vol.

27, No. 4, December.

Oliver S.J, 1998, The Performance of Concrete Topped Precast Hollow-core Flooring

Systems Reinforced with and without Dramix Steel Fibres under Simulated Seismic

Loading, Masters Thesis, Department of Civil Engineering, University of Canterbury,

Christchurch, Canterbury.

Trowsdale J.F, 2004, Seismic Performance of Hollow-core Seating Details Specified by

Amendment No. 3 to NZS 3101:1995, Research report submitted as part of third

professional year coursework, Department of Civil Engineering, Christchurch, New

Zealand.

Page 2-34
3. Section 3 Super-Assemblage Experimental Investigation

SECTION SUMMARY

Recent earthquakes and laboratory research has demonstrated the seismic vulnerability of

buildings constructed with precast prestressed concrete hollow-core floor systems. However,

follow-up research has shown that with simple detailing enhancements, significant

improvement in the seismic performance of hollow-core floor systems can be expected. The

present experimental research aims at validating several new detailing enhancements. Based

on previous research findings, the present super-assemblage experiment included the

following details: (i) a reinforced connection that rigidly ties the floor into the supporting

beam, (ii) an articulated topping slab cast onto a timber infill solution that bridges between

the hollow-core units and parallel longitudinal frame beams; (iii) a specially detailed plastic

hinge zone in the beams that support the floor, that concentrates damage away from the

hollow-core units; (iv) Grade 500E reinforcing steel used in the main frame elements; and (v)

mild steel deformed bars in the concrete topping in lieu of the customary welded wire mesh.

The full-scale structure is cyclically tested in both the longitudinal and transverse directions

to inter-storey drifts of 5%. Visual and instrumental observations from the experiment are

presented and discussed. The experimental observations demonstrate that with appropriate

detailing it is possible to avoid undesirable failure modes induced in the hollow-core flooring

and to shift all damage into the plastic hinge zones of the supporting moment resisting frame.

Recommendations for the forthcoming revision of the New Zealand Concrete Structures

Standard, NZS 3101, are made.

Page 3-1
3.1 INTRODUCTION

Research over the last decade at the Universities of Auckland and Canterbury has sought to

determine the seismic adequacy of precast buildings with hollow-core flooring systems in

New Zealand and to provide better solutions to preserve life safety. One of the major concerns

identified from both overseas earthquake events and experimental investigations was the

connection of these precast hollow-core units to the surrounding lateral load resisting system.

Under earthquake loading that induces significant storey lateral displacements, the potential

for the hollow-core units to lose their seating, leading to a possible partial or full collapse of

the floor, focused research into the connection of the floor units with the surrounding

perimeter beam. The collapse of the hollow-core units during testing by Matthews (2004) and

Matthews et al (2003) identified serious flaws for existing precast concrete frame structures

with hollow-core flooring structural systems. A continuation of this research (Lindsay, 2004

and Lindsay et al, 2004) demonstrated that enhanced reinforcement details could be

implemented into new structures that should result in improved behaviour in a seismic event.

This research project is a further continuation of previous work on existing structures

(Matthews, 2004, and Matthews et al 2003, 2004) and on new structures (Lindsay, 2004, and

Lindsay et al 2004).

The popularity and widespread use of precast concrete is recognised in New Zealand

design standards where there is specific reference to precast concrete flooring support

conditions. Amendment No. 3 to the current New Zealand Concrete Structures Standard NZS

3101:1995 provides two specific acceptable solution details for the connection of hollow-

core floor units to reinforced concrete frame supporting beams. The second solution specifies

a reinforced connection that rigidly ties the floor into the supporting beam, but to this point

remains untested in any large-scale three-dimensional experiments. The research reported

herein presents an experimental investigation of the effectiveness of this solution and its

Page 3-2
adequacy for inclusion into the upcoming revised New Zealand Concrete Code and use in

New Zealand construction practice.

Following a summary of the findings from previous research, this section initially

provides an overview into the experimental set-up. The construction of the test specimen and

the modifications made to the concrete moment resisting frame super-assemblage from the

previous test programs are discussed. The specific details of this series of experiments are

then provided including the hollow-core seating details, the lateral connection between the

hollow-core unit and the parallel perimeter beams as well as the diaphragm reinforcement.

The set-up of the experiment, including the material properties of the specimen, the

instrumentation and the displacement drift controlled loading regime used are then discussed.

The second part of this chapter provides the visual and instrumental observations of the

testing. The hysteretic behaviour and strengths in terms of load displacement relationships of

the system are shown. The results and overall effectiveness of the design changes and new

features are discussed and concluding remarks are made.

3.2 FINDINGS FROM PREVIOUS RESEARCH

Owing to the widespread use of hollow-core flooring in precast concrete construction, an

experimental program was initiated to examine the seismic adequacy of such systems in the

New Zealand building stock (Matthews, 2004; Matthews et al 2003, 2004). A two bay by one

bay concrete moment resisting frame was constructed to represent a one-floor corner segment

of a typical New Zealand 10-20 storey precast concrete frame building. General results from

that program showed that although the concrete frame performed relatively well, the semi-

rigidly supported hollow-core flooring system, behaved poorly and collapsed at a relatively

low level of seismic intensity. The main source of damage occurred due to the relative

rotation between the supporting beams and floor and the expected sliding action of the floor
Page 3-3
on the supports provided by the beams did not occur. A snapping action at the hollow-core-to-

support beam connection and in particular within the webs of the hollow-core led to the

eventual failure of the floor. Recommendations for future connection details were made to

ensure that any inelasticity and relative rotation was centred at the interface between the floor

and supporting beam thereby protecting the hollow-core units. The connection between the

floor and the frame spanning parallel to the hollow-core units also performed poorly as a

result of the displacement incompatibility between the flooring, which is designed to sag and

the perimeter beams, which deform in double curvature. The undesirable deformation

demands placed on the hollow-core caused internal web cracking, fracturing of the cold

drawn mesh that reinforces the diaphragm and ultimately, the failure of the lower half of the

first hollow-core unit next to the two bay perimeter beams. Analytical work provided a more

accurate means of determining the lateral strength capacity of the structure accounting for the

influence of the floor slab contributing to the strength of the frame. A theory was developed

to predict beam elongation (Matthews et al, 2004). A rainflow counting theory was

advanced and validated against the experimental results. This provided a suitable method to

calculate required seat widths for precast floor systems.

In response to the poor behaviour of the hollow-core flooring during the Matthews

(2004) experiment, a Technical Advisory Group on precast floors (TAG) was formed and new

and improved seated connection details were recommended. Proof-of-concept sub-assemblage

tests were carried out by Bull and Matthews (2003) under sponsorship from Precast NZ Inc.

One of the new beam to floor connection details, which featured a low-friction bearing strip

with a compressible backing board replacing the plastic bungs in the end cores of the units,

performed well and was consequently recommended for firstly full-scale testing, and

subsequent inclusion into a revised New Zealand Concrete Structures Standard.

Page 3-4
Lindsay (2004) repaired and reconstructed Matthews (2004) full-scale super-

assemblage specimen and then retested it. Whereas the Matthews (2004) experiment was

retrospective examination of past construction techniques (1985-2000), Lindsay (2004)

looked towards developing solutions for improved performance for future precast concrete

construction practice. The seating detail included a low friction bearing strip and

compressible backing board, and was implemented along with a connection isolating the first

hollow-core unit from the perimeter beams by way of a timber infill solution. A more

adequate connection was used which tied the central column into the floor plate. These

modifications produced significantly improved behaviour, and the super-assemblage

withstood up to inter-storey drifts up to +5%. Despite the good performance, the bearing strip

was observed to slide on the seat (the ledge cast along the beams that support the precast

hollow-core floor), the compressible backing board did not compress and hollow-core units

seated on or near the highly deformed plastic hinge zones of the supporting beams showed

cracking across the corners of the hollow-core units. The concrete of the seat spalled

considerably, and although the hollow-core was relatively undamaged, the units were

beginning to slide off the seat and down the face of the supporting beam. The infill detail

behaved well and proved to be a suitable flexible link isolating the beams and floor. However

premature fracture of the diaphragm ductile mesh produced a longitudinal tear within the

topping concrete along the interface between the timber infill and the first hollow-core unit.

Other research programs have looked at hollow-core connection detailing at a sub-

assemblage component level. Meija-McMaster and Park (1994), Herlihy et al (1995), Oliver

(1998) and Herlihy and Park (2000) tested various reinforced hollow-core to beam

connections. Tests were primarily featured pull-off tension loads, to determine the adequacy

of reinforcing to be able to carry the weight of the floor if a loss of seating occurs. Among the

results, the tests showed that plain round reinforcing could provide suitably ductile behaviour.

Page 3-5
By comparing the aforementioned sub-assemblage component research with the

failure modes witnessed during the Matthews (2004) experiment, it was realised that the

earlier work was fundamentally flawed in that the rotational component of deformation was

not included in the sub-assemblage experiments. Subsequently, testing by Bull and Matthews

(2003) and Liew (2004) included a cyclic vertical force into the testing set-up to create cyclic

relative rotations between the supporting beam and precast floor slab. The experiment

conducted by Liew (2004) showed that over-reinforced connections could be detrimental

towards performance. By forming a strong, rigid connection between the floor slab and

supporting beam, the relative rotational demand concentrated the damage at the relatively

weaker regions in the hollow-core floor slab away from the supporting beam. The brittle

nature of hollow-core meant that failure occurred at low imposed drift demands. It is therefore

important that there is a consideration of the influence of reinforcing the connection between

the hollow-core units and the supporting beams.

3.3 SUPER-ASSEMBLAGE SPECIMEN DESIGN DETAILS AND CONSTRUCTION

3.3.1 Hollow-core Seating Details

The seated connection detail used for this experiment is the second acceptable solution

detail given in Amendment No. 3 to NZS 3101:1995. The connection features two of the four

cores within the hollow-core units reinforced and filled with concrete, and is

diagrammatically shown in Figure 3-1(c). Deformed 12mm Grade 300 (D12) reinforcing is

used at 300mm centres for the starter bars, which is lapped with the same steel as the

diaphragm reinforcement. In the two reinforced cores, R16 Grade 300 bars (the R denotes

plain round bars) were placed close to the bottom of the cores, with both ends hooked to

provide sufficient anchorage in the beam and hollow-core.

Page 3-6
Reaction Pinned steel
frame tie beams

Unit 4
750x750
columns 300 series Unit 3
Transverse hollowcore
beams units Unit 2
475x750
N Unit 1 transverse beams

W E
750 wide 400x750
timber infill Plan longitudinal beams
S
6100 6100 6100
3500

Double acting Pinned fixed


roller bearings column
Front elevation Side elevation
(a) Details and geometry of the super-assemblage.
Starter bars from the
longitudinal beams Column
(extended into first hollowcore unit)

Hollowcore Units
Timber Infill

Transverse Beam
Hooked Bar
(in addition to the regular reinforcing) Longitudinal Beam

(b) Design modifications incorporated within the super-assemblage


D12 Starter Bars (Grade 300) Floor Unit

Supporting Beam
Low friction
bearing strip

Concrete Seat (ledge)

2*R16 (Grade 300) - 1 per filled core


10mm 15mm
Seat Transverse
50mm
Reinforcement Low Friction
bearing strip 75mm
75mm seat (west end)
50mm seat (east end)
(c) Hollow-core Seating Connection Detail (d) Close-up view of the seating
75mm Concrete Topping
D12 Starters extending 600mm
into the first hollow-core unit
(lapped with diaphragm reinforcing) Timber in-fill
750mm

Hollowcore Units
200mm*25mm Timber boards

Perimeter beam

(e) Lateral Articulated Infill Slab Connection Detail


Figure 3-1 Super-Assemblage building and hollow-core connection details.

Page 3-7
To prevent concrete entering the two non-filled cores of each hollow-core unit, a backing

board was used in place of the more conventional end plug. This was used to ensure that the

rotation of the floor units relative to the beam occurs at the critical section at the beam-to-

floor interface and that the relatively brittle hollow-core unit did not experience high

rotational demands. The backing board substitutes for the conventional use of end plugs,

which create a concrete stub part way into the cores, which, under relative rotations, causes

prying and splitting forces within the un-reinforced hollow-core webs. The hollow-core unit

was seated on a low friction bearing strip and the seat widths are 50mm and 75mm at the east

and west ends of the test structure respectively as shown in Figure 3-1(c) and (d). The code

Amendment prescribes a 75mm seating, even so, it was decided to investigate the effect of a

shorter seat width for cases when hollow-core units arrive on site shortened, due to shrinkage

and creep, indicative of real construction practice.

3.3.2 Lateral Connection and Diaphragm Reinforcement

The infill detail used in this experiment stems from the recommendations made by Lindsay

(2004) and is shown in Figure 3-1(e). A 75mm concrete topping is cast on a 750mm long by

25mm thick timber plank infill running between the first hollow-core unit and the longitudinal

beams. The starter bars from the perimeter beams extend 600mm into the topping above the

first hollow-core unit. The longer starter bars and use of conventional reinforcement instead

of cold drawn or ductile mesh was used as a means to increase the ductility capacity at the

infill-hollow-core interface and mitigate the risk of fracture from occurring, as observed

previously (Matthews, 2004; Matthews et al 2003, 2004; Lindsay, 2004; Lindsay et al, 2004).

For this experiment, Grade 300 D12 bars (the D designates bars with deformations) have

been used at 300mm centres in both directions. The use of D12 bars in the topping slab is

beneficial because they are the same as the starter bars used, and all lapping is between the

Page 3-8
same type of reinforcement. The current standards as proposed in the amendment to NZS

3101:1995, state that starter bars should extend to the larger of either 20% of the hollow-core

span, or the development length (ld) plus an additional 400mm, which in some cases can be a

considerable length. In this case, the starter bars and diaphragm bars are the same cross

sectional area and spacing, meaning effectively full-length starter bars with no curtailment.

The starter and lapped bars were strain gauged to investigate the optimal length that starter

bars should be. Also, two XD20 drag bars (where the X denotes Grade 500 bars), which

were successfully implemented into the testing by Lindsay (2004), were cast through the

topping over the flooring and the central column to tie back the column to the floor.

3.3.3 Super-Assemblage Construction

The full-scale super-assembly specimen, shown in Figure 3-1(a), was a two-bay by one-bay

structure designed as a lower storey corner section of a multi-storey precast concrete moment

resisting frame building. The pretensioned flooring system runs parallel to the longitudinal

beams (east-west), past the central column, and are seated on the transverse beams. The frame

dimensions were identical to the original Matthews (2004) supper-assemblage to maintain the

same loading set-up and to be able compare results. The columns were 750mm square in

section, spaced at 6.1m centres with an inter-storey height of 3.5m. The longitudinal beams

were 750mm deep by 400mm wide, while the transverse seat beams were 750mm deep by

475mm wide. The super-assemblage was constructed in a similar fashion as it would be done

on a construction site. A photograph of the frame during construction can be seen in Figure 3-

2. A full set of construction drawings and photographs can be found in Appendix B.

The original super-assemblage experiment (Matthews, 2004) was a retrospective look

at the structural details of existing buildings, whereas the second experiment (Lindsay, 2004)

along with the present investigation are aimed at providing new solutions for future

construction. With this in mind, it was decided to use Grade 500E seismic reinforcement

Page 3-9
throughout the frame, which is currently the most commonly used grade of steel in New

Zealand construction practice. The concrete frame was designed according to NZS

3101:1995, and is in keeping with the original Matthews frame strength and hierarchy of

strength. Following weak beam-strong column capacity design principles, the beam plastic

hinges were designed to the same strength as the original cases, however, using Grade 500E

steel. The beams required 6-XD25 bars, the exterior columns required 12-XD25 bars and the

centre column contained 24-XD 25 bars.

Figure 3-2 Photograph showing the precast concrete frame after it has been erected.

During previous testing, the hollow-core units seated on the potential plastic hinge zones of

the supporting beams suffered damage due to the high deformation occurring in these zones.

To avoid such deformation demand on the floor units, a hooked bar was placed within the

beam to force the hinge to occur close to the column face as shown in Figure 3-1(b). Another

negative feature revealed in the previous experiments was the spalling of the seat concrete. To

overcome this, transverse reinforcement was placed within the seat to tie the seat concrete

back into the beam, as shown in Figure 3-1(c). A second generation of low friction bearing

strip was used which featured longer teeth to both increase the friction resistance on the

underside of the strip, which is in contact with the ledge of the seat, and to negate the effects

of seat surface roughness.

Page 3-10
3.3.4 Material Properties

New Zealand manufactured Grade 500E seismic reinforcement was used throughout the

super-assemblage, with the exception of the diaphragm and hollow-core-to-frame connection,

which used Grade 300 seismic reinforcement. Table 3-1 presents a summary of the tensile test

information while a full set of stress strain curves are presented in Appendix C.

Table 3-1 Reinforcing steel tensile test material properties.


Tensile Test Strength and Strain Characteristics
Location Reinforcing fy (MPa) fsu (MPa) Es (GPa) Esh (GPa) Hsh Hsu
Column & Beam Longitudional Steel YD25 536 675 214 3.27 0.0015 0.0128

Column & Beam Transverse Steel YR12 530 714 190 - 0.0005 0.0152
Support Connection R16 336 469 186 4.48 0.0031 0.0200
Starter Bars and Diaphragm D12 307 447 181 3.65 0.0029 0.0218
Drag Bars YD20 580 724 246 3.82 0.0012 0.0156

Ready mixed concrete with a specified target compression strength of 30MPa, a maximum

aggregate size of 19mm and a slump of 120mm was used to cast the precast specimen

components. The topping slab was a 30MPa pump mix with a maximum aggregate size of

13mm. Table 3-2 shows the 7-day, 28-day and test day cylinder compressive strengths

measured from an average of three 200mm long by 100mm diameter cylinders.

Table 3-2 Concrete compressive strength material properties.

Compression Test Strengths (MPa)


Location 7 Day 28 Day Test Day
Precast Frame Units 1 26.7 37.8 45.3
Precast Frame Units 2 21.2 30.1 41.2
Precast Frame Units 3 25.1 32.0 35.2
Splice Half Beams 31.8 - 46.0
Topping Slab & Half Beams 31.0 - 40.6
Grout: Column Reinforcing - - > 60

Page 3-11
3.4 EXPERIMENTAL SET-UP

The testing of the super-assemblage was completed in four Phases: (i) longitudinal loading;

(ii) transverse loading, (iii) re-loading in the longitudinal direction, and (iv) re-loading in the

transverse direction. For Phase 1 and 3 loading, the set-up shown in Figure 3-3(a) illustrates

the location of the secondary load frames and primary load frames, which were attached to the

south frame of the super-assemblage. In each bay there were two actuators in opposing

diagonal directions: one that is displacement controlled, the other force controlled. Actuators

attached to the bottom of the north columns were displacement controlled and used to

maintain the same inclinations as the south columns and thereby minimise the torsion in the

transverse beams. Figure 3-3(c) shows an elevation of the loading setup for the south centre

column. For the previous experiments by Matthews (2004) and Lindsay (2004) this was the

setup for the transverse loading only, but this setup remained for the entire duration of testing

for this experiment. These actuators, by applying equal and opposite forces were able to keep

the centre column at a suitable transverse (north south) inclination during the longitudinal

(Phase 1 and 3) loading again minimising the torsion of the longitudinal beams. During the

transverse Phase 2 loading, these actuators ensured that the central column was at the same

inclination as the corner columns. The configuration shown in Figure 3-3(b) shows the

transverse set-up that was attached to the east and west frames (as shown) to allow loading in

the orthogonal transverse direction.

3.4.1 Instrumentation

Due to the size of this test specimen, an extensive array of instrumentation was required to

acquire all the experimental data. A custom designed program and data acquisition system

developed at the University of Canterbury were used to record and save the data from all of

the various types of instruments at each increment of load. Similarly, a control program was

Page 3-12
developed, in house, to actively control each of the actuators to load the structure in a

displacement-controlled manner. For a full account of the control system and algorithms

utilised for the loading, refer to Matthews (2004).

Diagrams showing the instrumentation can be seen in Figure 3-4. The load was

monitored by load cells that were attached to each of the eight hydraulic actuators used during

the experiment. Similarly, rotary potentiometers were able to monitor the actuator

displacements. Each leg of the secondary load frames was strain gauged in a full-bridge

configuration allowing the load to be monitored and the redistribution of load through the

system to be followed.

A significant portion of the instrumentation was to measure the displacements of the

structure through inclinometers, linear potentiometers and displacement transducers. The drift

displacement controlled nature of the experiment was achieved by using inclinometers to

measure the inter-storey drift of the structure or inclination of the columns. Sonic

displacement transducers were used at the base of the columns to accurately measure the

column sliding movement. Linear potentiometers of different sizes were placed at numerous

positions of the building to measure any large displacements, such as the flexural, shear and

elongation deformations of the plastic hinges, movement of the hollow-core units and relative

movement of the timber infill. The floor diaphragm was also extensively instrumented by a

grid pattern of Demec points, where the distance measured between the points allow strain

patterns of the entire floor to be generated. Vertical levelling was undertaken at discrete points

on the floor to obtain a contour map of the vertical displacements of the diaphragm as well as

provided an absolute measure of the vertical displacement of the whole structure. Finally,

starter bar and hairpin reinforcement for the support connections, starter bars passing over

the infill and the bars tying the central column into the floor were also been strain gauged to

monitor the level of load or inelasticity that the reinforcement were subjected to.

Page 3-13
(a) Load frame and hydraulic actuator set-up for Phase I and 3 in-plane.
Secondary loading frame

Ac
tua
tor
D

E
ator
tu
Ac
Self equilibrating loading frame

(b) West end elevation of the load frame and hydraulic actuator set-up for Phase 2 out-of-plane
testing.
Hydraulic actuator (Ram H)

Back Steel Tie Frame Columns

Reaction frame

Hydraulic actuator (Ram G)

Reaction frame

Linear bearings

(c) Side Elevation view of the central column load frame and loading set-up.

Figure 3-3 Loading set-up for the three loading Phases.

Page 3-14
Strain Gauge Location Strain Gauge Locations
300 600 600 600
Timber Infill
3
T 1 2 4 5

B 1 2
Hollowcore Units 50mm pots
30mm pots measuring the
horizontal pull-off and vertical Potentiometers measuring the
drop of the floor units relative relative vertical displacement
to the supporting beam between the floor diaphragm
Low Friction bearing strip and perimeter beams
on the 50 or 75mm seat

(a) Hollowcore support potentiometers and (b) Timber infill potentiometers (16 sets) and
strain gauge instrumentation. strain gauge instrumentation.

Pots to measure half


beam elonation

Front elevation Displacement Transducers - One


in either direction on front columns

Displacement Transducers - One


Pinned fixed in in-plane direction on back columns
column

Side elevation

(c) Location of displacement transducers and Inclinometers.


Potentiometers (top and bottom)
measuring the fixed-end rotation Potentiometers measuring
and curvature of the beam plastic longitudinal beam bar slip
hinges through the beam-column joint

Beam Plastic Hinge


Beam-Column Joint Potentiometer
Potentiometer instrumentation
instrumentation - measuring
- measuring hinge shear and
joint shear deformation
flexural deformation

(d) Location of beam plastic hinge and beam column joint potentiometers.

Figure 3-4 Instrumentation used for the super-assemblage testing.

Page 3-15
3.4.2 Loading Protocol

Experimental loading protocols at the University of Canterbury have commonly been based

on cyclic loading at increasing ductilities (Park, 1989). For the Matthews (2004) experimental

investigation into the seismic performance of existing structures a new loading pattern was

developed. The aim of the displacement loading regime was that it would be indicative of

what a similar structure would experience during a real earthquake event. The loading

protocol was based on studies into probable drift demands of a range of precast concrete

frame buildings subjected to time history analyses under a suite of different earthquake

records.

For the Lindsay (2004) experiment, the experiment was centred on investigating the

seismic behaviour of newly designed or constructed structures. Therefore, a more

conservative assessment of the seismic capacity was sought and hence a higher displacement

demand was necessary. This experiment also investigates the performance of a new solution,

and is utilising the same loading protocol as what was used during the Lindsay (2004) tests

offering a direct comparison of the results. The experiment consists of four phases comprising

of: (i) longitudinal loading of two completely reversed cycles of 0.5%, 1% and 2%; (ii)

transverse loading of two completely reversed cycles of 0.5%, 1%, 2% and 3%; (iii)

longitudinal re-loading of an initial cycle to 2% followed by two completely reversed cycles

of 3%, 4% and 5%; and (iv) transverse re-loading of one complete reversed cycle of 3%,

and then two completely reversed cycles of drift amplitudes of 4% and 5%. The Matthews

(2004) or Lindsay (2004) experiments did not conduct Phase 4. Therefore, for comparative

purposes, only the first three phases of results are reported herein.

Page 3-16
3.5 EXPERIMENTAL OBSERVATIONS

3.5.1 Phase 1: Longitudinal Loading

Phase 1 longitudinal loading included two cycles at increasing drift amplitudes to 2%. Some

photographs of the key damage are shown in Figure 3-5. Throughout the descriptions of the

test results, a positive inter-storey drift refers to when the tops of the columns move in an east

direction and the bottoms of the columns move west. From the early stages of testing,

diagonal torsional cracks appeared at the ends of the transverse beams and continued to

extend and widen throughout this phase of testing. Figure 3-5(b) shows early stages of torsion

with X type cracking. Cracks, propagating diagonally outwards from the longitudinal beams

across the infill slab first appeared at +0.25% and continued to extend into the topping

concrete of the first hollow-core unit. Figure 3-5(c) shows the tendency of the crack pattern to

arch towards the south central column. At +1%, a longitudinal crack was observed between

the first and second hollow-core unit, and after the 2% cycles this had extended to be around

6m long. Damage in the plastic hinge zones was largely restricted to one major crack at the

beam to column interface, and another significant crack around 300mm from the column.

Some instances of spalling of seat cover concrete were evident at +1% and after the 2%

cycles, the spalling had not extended but had worsened in a few areas as can be seen in Figure

3-5(d). A single continuity crack formed along the beam to floor interfaces, as shown in

Figure 3-5(a), hairline soffit cracks were first noticed underneath the a filled core at the east

end of Unit 1 and then a web crack was observed at +2% propagating at 45 degrees from the

seat to the topping. Beam elongation was illustrated both with the residual crack openings and

by the sliding of the floor units off the supporting beams (which could be seen by the

exposure of unpainted soffit of the hollow-core floor units). Results showed that the bearing

strips worked as expected. The residual drift after the 2% cycles was around 1.1% and the

structure had suffered moderate but possibly repairable damage.

Page 3-17
(a) Continuity crack forming at the east end at (b) Torsional cracking of the North West
+2%. Transverse Beam at 1%

(c) Cracking over the infill and floor of the (d) Removed spall concrete at east end and
west bay (looking towards south centre exposure of the bearing strip after Phase I
column) after Phase I testing. testing.

Figure 3-5 Damage to the test super-assemblage during and after Phase I longitudinal
loading.

Page 3-18
3.5.2 Phase 2: Transverse Loading

A selection of photographs showing the behaviour of the test specimen under Phase 2

transverse loading is shown in Figure 3-6 where the convention for postive drift is when the

tops of the columns move north and the bottoms of the columns move south. The early stages

of testing exhibited little new damage, and behaviour was characterised by the pre-existing

cracks (caused by Phase 1 loading) opening wider. In a similar fashion to the longitudinal

beams, the rotation experienced by the transverse beams was concentrated at or near the

column face rather than being distributed over a conventional plastic hinge zone length.

Figure 3-6(a) and (b) show the damage in these zones, where crack widths were

approximately 15mm at +3% drift. There was little new damage to the floor and topping slab

in general, although elongation of the transverse frames was clearly apparent with large

openings between the column and topping slab at the top of the beams under negative

moments. At +3% drift, the opening was around 15mm at the top of the transverse beams at

the north ends and the topping slab was pulling away from the south corner columns by

approximately 25mm at 3%. Vertical deformations of the floor also became apparent with

the floors dropping away from the column, which could also be seen from the shear

deformation along of the large cracks at the ends of the transverse beams. The large

deformations occurring at the northern end of the east transverse beam, due to shear and

flexure, resulted in a crack to propagate from the seat through the northernmost filled core of

Unit 4 and into the topping slab, as shown in Figure 3-6(c). This damage did not worsen

during the rest of the testing. After the completion of the Phase 2 transverse loading to 3%

the residual drifts were approximately 1.6% and the structure had suffered moderate damage

but was still in a repairable state.

Page 3-19
(a) Damage to the south end of (b) Damage to the north end (c) Hollowcore web cracking
the west transverse beam at of the west transverse beam at and damage at the north east
+3%. +3%. corner after Phase 2 testing.
Figure 3-6 Notable damage from Phase 2 transverse testing.

3.5.3 Phase 3: Longitudinal Re-Loading

Figure 3-7 shows some key photographs of the Phase 3 longitudinal re-loading of the test

specimen. During the initial stages of loading below 2% no additional significant damage

occurred. At +3% drift, several of the infill slab cracks extended over the hollow-core units

and small vertical displacements were first noticed (~1mm) at these cracks. A 1m long soffit

crack, 1-2mm wide, appeared at the west end of Unit 1 underneath one of the filled cores.

However no more damage to the hollow-core or beam seats of the supporting beams was

witnessed throughout the remaining testing. Upon loading towards 3% drifts, crushing of the

top concrete of the south centre column occurred at 2.4%, which can be seen in Figure 3-

7(d) and was matched by a small load drop off. It was also apparent to see the torsional nature

of the transverse beams. Whilst the south frame and columns were inclined the transverse

beams appeared to remain near vertical, acting to minimise the relative rotation imposed on

the hollow-core-to-beam seating connections. Cracking at the ends of the transverse beams

Page 3-20
showed between 2mm and 5mm of outwards movement and evidence of torsional hinging. It

was expected that the transverse beams would rotate rigidly with the corner columns with

slight inclination changes towards the midspan due to the eccentric loading of the flooring. At

4% drifts, significant amounts of concrete had become loose and fallen from the plastic

hinge zones at the ends of the south, east and west beams, and several of the reinforcing bars

became visible at the bottom of the southwest hinge and top of the southeast hinge of the

south beams. Although the structure was still in a stable condition, the damage in some areas

became irreparable and major components would need to be replaced for further structural

use. The phenomenon of beam elongation was again illustrated on several instances. Figure 3-

7(c) shows the gap opening of around 25mm wide at the southwest column, and Figure 3-7(e)

demonstrates by way of the unpainted floor, the amount of sliding of the floor units off the

supporting beams.

Prior to the first cycle to 4% drift, buckling of the compression bars at the bottom of

the southwest beam was observed, as illustrated in Figure 3-7(f) and at the top of the

southeast beam. On the accompanying cycles with an opposite bending moment, the bars did

not straighten completely in tension. During the final cycle of loading to 5% drift, at 1.14%

and then at 0.36% unloading from 5%, the inside and outside top reinforcing bars

respectively at the southeast beam hinge, fractured as shown in Figures 3-7(g) and (h).

Although the load-carrying capacity of the structural system was jeopardised, collapse

prevention was maintained; all but two reinforcing bars ensured that the frame remained

stable and the lack of damage to the hollow-core units and seating support mitigated any

major life safety concerns.

Page 3-21
(a) Super-assemblage specimen at +5% inter- (b) Damage to southeast plastic hinge at +5%
storey drift. of the south frame.

Direction of
movement

(c) Top view: crack opening (d) Top view: (e) Evidence of elongation and the
(~25mm) at between the southwest concrete crushing at floor units sliding on the bearings.
column and floor topping slab. south centre (seen as the unpainted soffit, Unit 1
column. west end)

(f) Longitudinal beam bar buckling (g) Longitudinal (h) Longitudinal beam bar fracture
(southwest beam bottom, south beam bar fracture at 0.36% (southeast beam top
frame). at 1.14% inside bar, south frame).
(southeast beam top
outside bar)

(i) Infill and floor damage after testing.


Figure 3-7 Damage during and after Phase 3 longitudinal reloading.

Page 3-22
3.6 HYSTERETIC PERFORMANCE AND SYSTEM STRENGTHS

Figure 3-8 presents the lateral base shear force displacement (drift) hysteretic performance of

the super-assemblage for each of the three loading phases. Figure 3-8(a) shows the behaviour

for Phase I loading, where a maximum positive and negative base shear capacities of 1460kN

and 1380kN was observed. These maxima both occurred on the first cycle to 2%, although

it must be noted that there was only a small reduction in base shear capacity during the second

cycle.

The hysteresis loop for Phase 1 loading exhibits a small reduction of stiffness at low

inter-storey drifts. This could be explained by the opening and closing of cracks and the

general slop associated with the pins within the load frames and actuators. Phase 2

conversely shows no premature deterioration of strength or stiffness and that the longitudinal

Phase 1 had little effect on the structural behaviour. The maximum positive and maximum

negative base shear forces were 755kN and 801kN respectively which both occurred on the

first cycle to 3%. There was no loss is base shear capacity during the second 3% cycle. The

non-symmetrical layout of the super-assemblage, with the tie frame at the north end can help

to explain the differences between the positive and negative base shear capacity values. Under

negative drift cycles, the south beam starter bars that pass over the infill tying the beam into

the floor are activated under negative bending moments. A crack extending along this beam-

infill interface opened up during testing to negative drifts, and was 8mm wide near the

columns and 1mm wide at midspan of the south beams. Under positive drifts there are no

starter bars along the northern edge of the floor to be activated, and hence a slightly smaller

base shear capacity is observed.

Page 3-23
2000

Base Shear
Force (kN)
1500
W E
1000

500
Inter-Storey Drift (%)
0
-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5
-500

-1000 W E
-1500

-2000

(a) Hysteresis behaviour for Phase 1 longitudinal loading.

1000
Base Shear
Force (kN)

S N 800

600

400

200
Inter-Storey Drift (%)
0
-4 -3 -2 -1 0 1 2 3 4
-200

-400
S N
-600 Pre Torsion
-800 Post Torsion

-1000

(b) Hysteresis behaviour for Phase 2 transverse loading.

2000
Base Shear
Force (kN)

W E 1500

1000

500
Inter-Storey Drift (%)
0
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
-500
W E
-1000
Phase 3
-1500 Phase 1

-2000

(c) Hysteresis behaviour for Phase 3 longitudinal re-loading.


Figure 3-8 Base Shear Force Inter-storey drift Hysteresis behaviour for the three
phases of testing.

Page 3-24
The hysteretic performance of the super-assemblage during Phase 3 longitudinal loading is

shown in Figure 3-8(c). The maximum positive base shear was 1583kN and the maximum

negative base shear force was recorded at 1536kN which both occurred during the first cycle

to 3%. Noticeable reductions in the base shear capacity were observed between the first and

second cycles to 4% and again to 5% - approximately a 14% decrease in base shear

capacity. Sudden drop-offs in load due to crushing of the topping concrete either side of the

south central column can be seen in Figure 3-8(c) firstly at 2.4% and at +3.5%. The

overlying hysteretic behaviour from Phase 1 shows that the damage from the transverse Phase

2 loading did not have a detrimental effect, and the maximum positive and negative base

shear forces were observed during Phase 3 testing. Subsequent longitudinal testing to high

drifts would, due to the longitudinal beam reinforcing bars fracturing (on the second cycle,

unloading from 5%), result in a reduction in the base shear capacity of the structure.

3.7 DISCUSSION OF RESULTS AND EFFECTIVENESS OF DESIGN CHANGES

3.7.1 Transverse Beam Torsion

Perhaps the most significant feature of the present experiment was the extent of torsional

twisting evident, under longitudinal loading, of the two transverse floor-supporting beams.

Five inclinometers were installed on the western transverse beam to gauge the amount of

torsion experienced. Figure 3-9(a) presents the torsion, in terms of percentage drift as a

function of time for Phase 1 and Phase 3 loading, and the true extent of the relative torsional

twist that occurred in the plastic hinge zones at the ends of the transverse beams is illustrated

in Figure 3-9(b). It can be seen that during early stages of testing, the rotation of the beam

followed that of the corner columns - rigid rotation which would be expected. In the latter

stages of Phase 1 the beam rotation was similar to the columns for positive drifts but less for

negative drifts. This can be explained by the eccentric loading of the flooring units rotating

Page 3-25
the beam in a positive direction, and torsional moments generated by the moment resisting

beam-floor connection reinforcement, as the torsional stiffness of the beam progressively

degrades. During testing, cracking was observed on the transverse beam hinges at early stages

and at -1%, X type torsion cracks were evident. The torsion behaviour for Phase 3 loading,

seen in Figure 3-9(a) showed interesting results and that the transverse Phase 3 loading had

had a marked effect. The results illustrate that the beams remained essentially upright

throughout the test. There is an apparent tendency towards positive drifts accounting for the

eccentric floor support. For example, at the first cycle to +4% the beam was at an average

+1% inclination, and the first cycle to 4%, the beam remained at around +0.1%. All five of

the inclinometers show similar behaviour. This indicates that the torsion occurred at the ends

of the beams in the plastic hinge zones. Similar behaviour and damage was observed at the

eastern end of the super-assemblage. Appendix D shows the location of the five west beam

inclinometers.

The presence of torsional hinges had important effects on the behaviour of the

structure. The first issue is that the torsion could reduce the strength of the flexural hinges

placing increased demand on the reinforcement and concrete. This could possibly jeopardise

the integrity of the structure. However, no major detrimental behaviour was observed. The

positive implications of the torsional hinging was that the relative rotation between the

flooring units and supporting beam, which is the primary cause of damage to the supporting

beam-to-floor connection, was effectively small and can help explain the improved

performance of the seated connection and lack of damage as compared to the floor tested by

Matthews (2004).

The torsional behaviour is quite different to what was observed during the previous

experiments conducted by Matthews (2004) and Lindsay (2004). Figure 3-10

diagrammatically illustrates the difference in behaviour between the experiments. It was

Page 3-26
expected that in a similar way to the previous tests, the torsion would be distributed over the

length of the beam rather than at torsional hinges at the beam ends. The current experiment

featured a lower amount of reinforcement (6XD25: U = 0.98% compared with 12D24: U =

1.8% of the earlier experiments) passing into the column, resulting in a reduced torsional

stiffness. However, the most likely cause of the torsional hinging is the effect of the

reinforcement connecting the supporting beam and hollow-core units. The rigid floor-to-beam

connection acted to keep the beam more upright, imposing an opposing torsional force as the

column and beam rotates. The floor-to-beam connection detail used in the Lindsay (2004)

experiment was in effect a pinned connection, and this lack of fixity of the connection detail

would mean that the additional torsion applied to the supporting beam would not be present.

A more detailed examination of the torsional behaviour observed is provided in Section 4.

6 Phase 3
5
Torsion Rotation (%)

4 Phase 1 loading
3 loading
2
1
0
-1
-2
-3
SW Column Inclination Inclinometer 1
-4 Inclinometer 2 Inclinometer 3
-5 Inclinometer 4 Inclinometer 5
-6

(a) West beam inclination during Phase 1 and 3 loading.


6 Phase 3
5
An gle of Twist (rad/m)

4 Phase 1 loading
3
2 loading
1
0
-1
-2
-3
-4
-5
-6
-7
-8

(b) Torsional twist of the south hinge of the west transverse beam
(radians/metre in plastic hinge zone)
Figure 3-9 Torsion rotational behaviour of the western transverse beam.

Page 3-27
Large rotation at the South-West
end of the beam Column

Small rotation at the


centre of the beam

Torsion distributed over


length of beam

Eccentric weight of
floor units

(a) Behaviour observed during Matthews (2004) and Lindsay (2004)


experiments (adapted from Matthews (2004))

South-West Large rotation at the South-West


Column end of the beam Column

3D effects of beam elongation -


front frame elongates and pulls beam
ends away from floor, but out-of-plane
bending and strong floor-beam
connections means small openings at
beam centre

Torsion distributed over short


N length of the beam ends

Small rotation at the


centre of the beam

(b) Behaviour observed during the current experiment,


under positive drift (left) and negative drift (right)

(c) Damage and evidence of (d) Torsion of the east (e) Torsion of the east
torsional hinging (northwest transverse beam at +5%. transverse beam at 5%.
connection).
Figure 3-10 Torsional behaviour and damage of the transverse beams.

Page 3-28
3.7.2 Hollow-core-to-Supporting Beam Connection

Beam torsion played a significant role in the performance of the super-assemblage. While

torsion individually may be considered a sign of adverse behaviour, it had a beneficial effect

on mitigating the damage within the hollow-core-to-supporting beam connection. Figure 3-9

shows the difference between the rotation of the columns and the west transverse supporting

beam. The low rotations measured by the inclinometers indicated that the transverse beams

were remaining close to vertical, and the hollow-core floor slabs horizontal. Therefore, the

relative rotations between the floor slab and supporting beam was minimal.

The connection detail performed admirably and the super-assemblage structure was

able to sustain inter-storey drifts up to 5% with damage restricted to the support beam alone

and not the floor. At the conclusion of testing, there was minor diagonal web cracking at the

eastern end, both at the south and north sides of Units 1 and 4 respectively. A camera that

could be placed into four of the hollow cores showed that there was no observable internal

cracking of those particular cores. Soffit cracking was observed under several of the filled

reinforced cores at both ends of the structure. However these cracks appear to have had little

effect on the performance. The single crack that formed along the support beam-to-floor

interfaces on the topping concrete and lack of any other cracking in the adjacent floor

confirmed the objective to concentrate the relative rotation on this plane and restrict the

rotational demand on the floor units. The bearing strip was also adequate and allowed the

floor units to slide relative to the supporting beam whilst staying on the seat. At the 4%

peaks, around 16mm of pull off was recorded and could be seen underneath the structure,

while the bearing strip on the seat remained in place. The improved friction resistance and

better grip associated with the newer generation of bearing strip seems to have improved the

performance of the bearing strip. The minimal amount of seat spalling was also a positive

feature of the experiment. It is difficult to pinpoint the exact reasons why there was little

Page 3-29
spalling, but aside from the torsion of the supporting beams reducing the amount of rotation

experienced by the supporting beams, the bearing strips evened out the surface irregularities

of the seat concrete and the presence of transverse reinforcing tying the seat concrete back

into beam would have helped.

The performance of the seated connection detail showed very similar performance

characteristics to the same connection detail tested under two dimensional sub-assemblage

conditions outlined in Section 2. The close similarity of observed performance between the

two-dimensional and three-dimensional experiments, illustrates the usefulness of two-

dimensional tests that are much easier to undertake in terms of time, cost and resources.

It must also be noted that there was no discernible difference in the behaviour between

the 50mm and 75mm seating ends of the structure. Minor spalling amounted for a loss of the

50mm seat, however this was confined to a small area of around 300mm wide. Although good

behaviour of both ends shows that the connection detail can be utilised over smaller than code

specified seat widths (75mm), short seat widths should not be specified. The experiment was

undertaken in a controlled situation where the seating details were constructed as designed

and the seat widths were measured. Specification of a minimum of 75mm seat accounts for

the likely maximum drifts while maintaining a minimum bearing width for the construction

tolerances and shrinkage and creep of the hollow-core units that occur, and specification

should adhere to minimum seat widths of 75mm or more. Furthermore, the laboratory

investigation was not a real time dynamic test such that neither vertical accelerations nor

higher mode effects were applied. This further underlines the need to maintain 75mm bearing

widths.

3.7.3 Supporting Beam Detailing Enhancements

To overcome some of the deficiencies observed during the Matthews (2004) and Lindsay

(2004) experiments, detailing improvements including a hooked bar in the transverse

Page 3-30
supporting beams, to promote column face hinging (see Figure 3-1), and seat transverse

reinforcing were implemented. The minor damage that occurred within the hollow-core units

at the southern corners of the structure point to the added hooked bar performing as designed

restricting the plastic hinge to the column face. The hinging itself was confined to a small

area, approximately within half the beam depth from the column face for all of the plastic

hinges within the structure, so it is difficult to say that the hooked bar was solely responsible.

However, the unit seated on the northeastern corner plastic hinge did experience web

cracking, which can be seen in Figure 3-6(c), which further supports the contention that it is

preferable to seat hollow-core floor units away from areas of high deformation within

supporting beams, and isolation infill-type details should be used over plastic hinge zones.

The presence of the second generation bearing strip possibly contributed to a reduction of

stress concentrations along the seat, thereby reducing the spalling (as compared to Lindsay

(2004)).

3.7.4 Timber Infill Connection and Diaphragm Reinforcing

The inclusion of the timber infill connection isolating the longitudinal beams from the floor

system performed moderately well during Lindsays (2004) experiment. However, that

experiment demonstrated the disadvantage of using relatively brittle welded wire mesh. In the

present experiment the provision for longer starter bars and use of conventional reinforcement

(instead of ductile mesh) proved to be successful. The crack pattern extended through the first

floor unit and across the second floor unit indicating that the longer starter bars were more

suitable and able to distribute the loads over a larger area. The use of conventional

reinforcement within the diaphragm also proved to be successful. Longer starter bars

throughout the topping concrete ensured that no major cracks appeared and therefore the

diaphragm reinforcement was not highly strained therefore avoiding failure.

Page 3-31
3.7.5 Displacement Incompatibility

Figure 3-11 presents the displacement incompatibilities between the perimeter longitudinal

beams and the first hollow-core floor unit at peak drift amplitudes under Phase 1 and 3

loading. The displacements are plotted along the length of the structure from the west end to

the east end and were derived from levelling carried out on the floor. The results shown

indicate the amount of relative vertical displacements between the floor units (Unit 1) and

south frame beams.

Figure 3-11(a) shows the vertical displacements for Phase 1 loading. It can be seen

that there was a relatively symmetrical relationship between the positive and negative drifts.

The maximum incompatibility was approximately -10.5mm at +2% drift on the west side of

the south centre column meaning that the floor had dropped relative to the beam. Figure

3-11(c) illustrates a similar trend but distinct shift in the negative direction indicates how

much the floor dropped following transverse loading. The damage incurred by the transverse

beam hinges was associated with a shearing displacement that lowered the beam and flooring

together. The effect of the transverse loading is also shown in Figure 3-11(b) where the 2%

cycle peak incompatibilities are compared for both loading phases. The crossover point that

occurred close to the centre of the structure (6100mm) shows the floors had dropped

approximately 8mm. The maximum incompatibility evident shows a 33mm drop of the floor

relative to the beam at +4% drift. This corresponds to a 4.4% slope over the infill in a north

south sense. Due to the lack of observed damage or tearing of the topping concrete, it can be

deduced that such slopes could be accommodated by the infill topping concrete and starter

bars. Further displacement incompatibility results from the experiment, and accompanying

computer modelling and analysis is located in Section 4.

Page 3-32
10
Position along fram e (m m )
5

0
Displacement (mm)

0 2000 4000 6000 8000 10000 12000


-5

-10

-15

-20 Phase 1: 1%
Phase 1: -1%
-25 Phase 1: 2%
-30 Phase 1: -2%

-35

(a) Relative vertical displacement incompatibilities during Phase 1 loading (displacement of the
floor with respect to the longitudinal beams).
10

5
Pos ition along fram e (m m )
0
Displacement (mm)

0 2000 4000 6000 8000 10000 12000


-5

-10

-15

-20
Phase 3: 2%
-25 Phase 3: -2%
Phase 1: 2%
-30
Phase 1: -2%
-35

(b) Comparison of the 2% cycle displacement incompatibilities between Phase 1 and Phase 3.

10

5
Position along fram e (m m )
0
Displacement (mm)

0 2000 4000 6000 8000 10000 12000


-5
-10

-15
Phase 3: 2%
-20 Phase 3: -2%
Phase 3: 3%
-25 Phase 3: -3%
Phase 3: 4%
-30
Phase 3: -4%
-35
(c) Relative vertical displacement incompatibilities during Phase 3 loading (displacement of the
floor with respect to the longitudinal beams).

Figure 3-11 Vertical displacement incompatibilities between the floor and longitudinal
beams during Phase 3 loading. (Note a negative displacement means that the floor has
dropped relative to the beams)

Page 3-33
3.8 CONCLUSIONS

From the experimental investigation that has been presented in this section, the following

conclusions are made:

1. A two-bay by one-bay reinforced concrete precast frame with one-way hollow-core

flooring super-assembly was designed and constructed in accordance with

NZS3101:1995 and included the Amendment No. 3 rigid connection hollow-core seating

detail. The specimen performed well up to inter-storey drifts of 5% when the

longitudinal reinforcing bars in the beam fractured due to low cycle fatigue. The floor

system sustained only minor damage, which in no way impaired life safety. The onset of

damage in the plastic hinge zones and hollow-core-to-supporting beam connections

occurred at +1% and irreparable damage to the plastic hinges zones of the south beams

occurred at +4% during longitudinal re-loading.

2. The rigid floor-to-supporting beam connection behaved well up to structural inter-storey

drifts of 5%, where damage to the hollow-core units and seated support was minimal.

Previously, the details investigated herein were untested, and for that reason Amendment

No. 3 to the current New Zealand Structures Standard NZS 3101:1995 limited inter-

storey drifts of this class of construction to 1.2%. In light of the good performance, this

restriction should be removed.

3. The articulated timber infill slab connecting the lateral load resisting system longitudinal

beams with the floor slab performed well with no major cracking and tearing

experienced. The use of longer starter bars and conventional reinforcement in place of

the cold drawn and ductile mesh ensured that any displacement incompatibility was

accommodated and diaphragm action was maintained. The performance also verifies the

details proposed by Amendment No. 3 to NZS3101:1995.

Page 3-34
4. Three-dimensional effects of torsional hinging have arisen throughout the course of this

investigation. A rigid beam-to-floor connection acted to keep the floor horizontal and

beams upright forcing torsion of the transverse beams to occur over a limited zone in the

plastic hinges at the ends of the transverse beams under high drift rotations of the super-

assemblage. More research and analysis of torsion within precast concrete structures is

required.

5. Precast concrete moment resisting frames designed and constructed with Grade 500E

reinforcement can be expected to perform well under substantial seismic displacements.

Plastic hinge zones appear to be less than half the depth of the beam long, and most

rotation is centred at large cracks at the beam-column interface. More research needs to

be carried out utilising various reinforcement ratios and investigating the buckling and

low cycle fatigue characteristics of Grade 500E reinforcement within three-dimensional

concrete frames.

3.9 REFERENCES

Bull D.K, Matthews J.G, 2003, Proof of Concept Tests for Hollow-core Floor Unit

Connections, Precast NZ report, Feb 2003.

Herlihy M.D, Park R and Bull D.K, 1995, Precast Concrete Floor Unit Support and

Continuity, Research Report 93/103, Department of Civil Engineering, University of

Canterbury, Christchurch, New Zealand.

Herlihy M.D and Park R 2000, Precast concrete floor support and diaphragm action,

Research Report 2000/13, Department of Civil Engineering, University of Canterbury,

Christchurch, New Zealand.

Page 3-35
Liew H.Y, 2004, Performance of Hollow-core Floor Seating Connection Details, Masters

Thesis, Department of Civil Engineering, University of Canterbury, Christchurch,

New Zealand.

Lindsay R.A, 2004, Experiments on the seismic performance of hollow-core floor systems in

precast concrete building, Masters Thesis, Department of Civil Engineering,

University of Canterbury, Christchurch, New Zealand.

Lindsay R.A, Mander J.B, and Bull D.K, 2004a, Preliminary results from experiments on

hollow-core floor systems in precast concrete buildings, 2004 NZSEE Conference,

Rotorua, NZ, Paper 33, 8 pp.

Lindsay R.A, Mander J.B, and Bull D.K, 2004b, Experiments of the Seismic Performance of

Hollow-core Floor Systems in Precast concrete Buildings, 13th World conference on

Earthquake Engineering, Vancouver, Canada, CD ROM paper No. 585.

Matthews J.G, 2004, Hollow-core floor slab performance following a severe earthquake, PhD

Thesis, Department of Civil Engineering, University of Canterbury, Christchurch,

New Zealand.

Matthews J.G, Bull D.K, and Mander J.B, 2003a, Background to the Testing of a Precast

Concrete Hollowcore Floor Slab Building, 2003, Pacific Conference on Earthquake

Engineering, February, Christchurch, NZ, CD ROM paper 170.

Matthews J.G, Bull D.K, and Mander J.B, 2003b, Preliminary Experimental Results and

Seismic Performance Implications of Precast Floor Systems with Detailing and Load

Path Deficiencies, 2003 Pacific Conference on Earthquake Engineering, February,

Christchurch, NZ. CD ROM paper 077.

Matthews J.G, Mander J.B, Bull D.K, 2004, Prediction of beam elongation in structural

concrete members using a rainflow method, 2004 NZSEE Conference, Rotorua, NZ,

Paper 27, 8 pp.

Page 3-36
Meija-McMaster J.C and Park R, 1994, Tests on Special Reinforcement for the End Support

of Hollow-core Slabs, PCI Journal, Vol 39, No. 5, pp. 90-105, September-October.

NZS3101, 1995, Concrete Structures Standard, NZS3101, Parts 1 & 2, Standards New

Zealand, Wellington, New Zealand.

NZS3101, 2004, Amendment No. 3 to 1995 Standard (NZS3101), Standards New Zealand,

Wellington, New Zealand.

Oliver S.J, 1998, The Performance of Concrete Topped Precast Hollow-core Flooring

Systems Reinforced with and without Dramix Steel Fibres under Simulated Seismic

Loading, Masters Thesis, Department of Civil Engineering, University of Canterbury,

Christchurch, Canterbury.

Park R, 1989, Evaluation of ductility of structures and structural assemblages from

laboratory testing, Bulletin of the New Zealand National Society for Earthquake

Engineering, Vol.22, No. 3, September, pp 155-166.

Page 3-37
4. Section 4 Forensic Analysis of Failure and Behaviour Modes

SECTION SUMMARY

Present experimental research has sought to verify new detailing requirements for the use of

hollow-core flooring in precast concrete structures. A full-scale two bay-by-one bay concrete

frame structure with hollow-core flooring is cyclically tested to increasing lateral

displacements in both the longitudinal and transverse directions. This chapter provides a

forensic analysis of the present research experimental observations. Torsional hinging

behaviour that is observed due to a strong hollow-core floor connection with a weak

supporting beam is explained. A comparison is made between the calculated torsional

capacity and torsion demands. The state-of-the-art understanding of beam-to-floor

displacement incompatibility is also discussed. Experimentally observed displacement

incompatibilities are compared with those determined computationally. Examples of the

potential of modelling for future research and construction are provided. The overall base

shear capacity of the structure is predicted and is shown to agree well with experimental

results. To achieve this calculation, it is demonstrated that it is necessary to embody all

sources of strength enhancement, including a so-called bowstring effect whereby

compressive forces are transferred into beams running parallel and adjacent to the prestressed

concrete floor units. Finally, the present research is compared and contrasted with previous

similar experimental investigations. It is concluded that through subtle detailing

enhancements there can be improvements in seismic performance whereby a relatively fragile

hollow-core flooring system is transformed into a robust frame-floor system where all the

principal sources of damage are developed into the ductile frame.

Page 4-1
4.1 INTRODUCTION

Over the previous two decades, prestressed concrete hollow-core flooring systems have

become one of the preferred forms of precast construction in New Zealand. This research

described herein seeks to explain certain key experimental observations made in a lateral

loading test conducted on a two-bay by one-bay full-scale precast concrete super-assemblage,

which is shown in Figure 4-1.


Reaction Pinned steel
frame tie beams

Unit 4
750x750
columns 300 series Unit 3
Transverse hollowcore
beams units Unit 2
475x750
N Unit 1 transverse beams

W E
750 wide 400x750
timber infill Plan longitudinal beams
S
6100 6100 6100
3500

Double acting Pinned fixed


roller bearings column
Front elevation Side elevation

(a) Super-assemblage details.

(b) View of the floor of the super-assemblage. (c) Load frame set-up.
Figure 4-1 Super-assemblage test specimen details.

Page 4-2
The experiment is the third super-assemblage examination on precast concrete structures. The

first of these was conducted by Matthews (2004) on existing buildings with pre-2001

construction details. Matthewss work discovered that existing hollow-core flooring details

performed unsatisfactorily and can be expected to behave poorly in a design basis earthquake

(10% probability of occurrence in 50 years). As an upshot of the work by Matthews (2004),

Amendment No. 3 was fast-tracked into the New Zealand Concrete Design Standard (NZS

3101, 2004). In Amendment No. 3, two acceptable solutions are given for detailing hollow-

core-to-supporting beam connections. The first of these was tested by Lindsay (2004). This

current research has investigated the viability of the second acceptable solution. Where past

research has focused on lateral load resisting systems overlooking the effect of floor

diaphragms, the interaction between the two became a principal focus for these investigations.

One significant finding in the current strong floor-to-weak support beam experiment

was the extent of the plastic torsional response of the supporting beams. This section seeks to

provide a forensic analysis of these experimental results. Several other failure mode aspects of

the super-assemblage specimen are also examined in an analytical fashion.

The performance of the floor diaphragm is assessed with particular emphasis on

displacement incompatibility between the hollow-core flooring and the adjoining parallel

perimeter concrete frame. Previous experimental findings and analytical work on

displacement incompatibility is summarised and the experimental results are compared with

results from computer modelling. New design solutions to combat the problems associated

with displacement incompatibility and reducing floor slab damage and the potential benefits

for furthering this work are also outlined.

The lateral force-deformation capacity of the structural system is assessed using a

theoretical pushover analysis. Comparisons of the theoretical and experimental lateral force-

deformation capacity results are provided and reasons for the discrepancies are explained.

Page 4-3
Additional results obtained from the experiment are compared with existing theoretical

relationships where the beam elongation values are evaluated against the rainflow analysis

methodology proposed by Matthew et al (2004). Throughout this section, the current

experiment is compared and contrasted with respect to the preceding two investigations

(Matthews, 2004 and Lindsay, 2004) as a means to complete the overall investigation into the

seismic adequacy of hollow-core flooring supported by concrete frame structures.

4.2 TORSIONAL BEHAVIOUR OF THE TRANSVERSE SUPPORT BEAMS

The behaviour of the transverse beams from observations of the experiment demonstrated

interesting torsional behaviour. Figure 4-2 presents the torsional behaviour of the west beam

observed during the experiment. Figure 4-2(c) shows the amount of torsional twist that

occurred within the south hinge of the west transverse beam. The rotations were measured

from inclinometers spaced evenly along the west transverse beam, which can be seen in

Appendix D. The results show that torsional hinges formed at the ends of the beams and that

the middle of the beam remained vertical and the flooring horizontal for most of the testing.

The comparison between Phase 1 and Phase 3 shows that there was a large increase in the

torsional twisting at the ends of the transverse beam following Phase 2 transverse loading.

During Phase 2, the ends of the transverse beams formed flexural plastic hinges, and as a

result of the damage caused due to the high flexure and shear actions, the torsional stiffness of

the ends of the transverse beams was significantly reduced.

This sub-section examines the experimentally observed torsional behaviour from

Phase 1 and 3 longitudinal loading and compares the observations with calculated torsional

demands and capacities. The torsional behaviour of the transverse beam was not measured or

analysed during Phase 2 transverse loading, but it is explained that the damage sustained by

the transverse beams during transverse loading has a significant impact on the torsional

behaviour during Phase 3.

Page 4-4
South-West Large rotation at the South-West
Column end of the beam Column

3D effects of beam elongation -


front frame elongates and pulls beam
ends away from floor, but out-of-plane
bending and strong floor-beam
connections means small openings at
beam centre

Torsion distributed over short


N length of the beam ends

Small rotation at the


centre of the beam

(a) Schematic representation of the torsional behaviour


6
5
Torsion Rotation (%)

4
Phase 1 Loading Phase 3 Loading
3
2
1
0
-1
-2
-3
SW Column Inclination Inclinometer 1
-4 Inclinometer 2 Inclinometer 3
-5 Inclinometer 4 Inclinometer 5
-6
(b) West beam inclination during Phase 1 and 3 loading.
6
5
Angle of Twist (rad/m)

4
3 Phase 1 Loading Phase 3 Loading
2
1
0
-1
-2
-3
-4
-5
-6
-7
-8

(c) Torsional twist of the south hinge of the west transverse beam
(radians/metre in plastic hinge zone)
Figure 4-2 Observed torsional behaviour of the west transverse beam

Page 4-5
Figure 4-3(a) shows an idealised diagram of the hypothesised relationship between the

torsional strength (torsion moment) and torsional twist. In a similar way to the calculation of

the shear capacity of a reinforced concrete beam section, the torsional strength comprises of

strength contributions from the concrete, Tc, and the reinforcement, Ts. The ultimate torsional

moment strength, Tu, is a sum of the concrete and reinforcement contributions and is shown as

the upper bound torsional moment strength in Figure 4-3(a).

Torsion
Strength Monotonic torsion
capacity: Tu = Tc + Ts
Ultimate Torsion
Strength, Tu = T c + Ts

Tc
GKc r Cyclic torsion capacity:
Tu = Ts
Torsional Demand
Cracking Torsion = Tc

Reduced torsion capacity Ts


OTc
GK flexure, shear and cyclic
Stiff pre-cracking interaction
response

Torsion
Twist
Phase 3 response: soft post-
Phase 1 response: initial stiff pre-
cracking response due to damage
cracking response and small torsional
during Phase 2 transverse loading,
twist rotations
large torsional twist rotations

O = load path dependent parameter (O<1) that depends on the degree of previous loading and
flexure-shear-torsion interaction

(a) Schematic relationship between the torsion force and torsional twisting

South-West
Column

Torsional
Capacity of the
beams resisting
North-West the torsion at the
Column beam ends

Reinforced Torsional Demand


Connection under from the eccentric floor Reinforced
positive moments loading and reinforced Connection under
connection moment negative moments

(b) Schematic illustration of torsional the demand and capacity.

Figure 4-3 Idealised torsional behaviour of the west transverse beam

Page 4-6
The derivation of the torsional capacity of reinforced concrete sections is complex, and it is

difficult when considering interaction with flexure and shear actions and pre- and post-

cracking behaviour. Significant work has been undertaken by Hsu (1993) and Mitchell and

Collins (1991), particularly with prestressed structures, and Park and Paulay (1975) for

reinforced concrete structures. These sources provide examples for the calculation of torsional

demands and capacities for concrete structural elements and several aspects of this work have

been reflected in the New Zealand Concrete Structures Standard, NZS 3101: 1995.

For compatibility torsion, which is the case in this experiment, a reduction of the

torsional moment can occur due to the redistribution of internal forces when cracking first

occurs. NZS 3101:1995 provides an equation for calculating the cracking torsional moment

for a section, and is the same as the formulation given by Mitchell and Collins (1991). The

cracking torsion, or maximum design torsional moment can be calculated by:

f c' Acp2
Tc I (4.1)
3 p cp

where Tc = design torsional moment force at section at the ultimate limit state; or cracking

torsion; I = strength reduction factor; fc = concrete compressive strength; Acp = area enclosed

by the outside perimeter of the concrete cross-section; and pcp = outside perimeter of the

concrete cross section.

After diagonal cracking has occurred, a truss mechanism of vertical transverse

reinforcement ties and diagonal concrete compression struts form, and the contribution to the

torsional resistance from the reinforcement can be calculated. Park and Paulay (1975) provide

an expression for the reinforcement contribution to torsion capacity as a function of the

transverse and longitudinal reinforcement, which is shown as Equation 4.2. The equivalent

design expression for the nominal torsional strength from NZS 3101:1995 relates to the

Page 4-7
amount of transverse reinforcing and stems from the work presented by Park and Paulay

(1975).

At f y Al f ly
Ts 2 A0 (4.2)
s P0

where Ao = area enclosed by the connecting lines between the centres of the longitudinal

reinforcement; At = area of a transverse reinforcement stirrup leg; fy = yield strength of the

transverse reinforcement; s = transverse reinforcement spacing; Al = total area of the

longitudinal reinforcement; fly = yield strength of the longitudinal reinforcement; and Po =

perimeter formed by area Ao.

Prior to torsional cracking occurring or the torsional strength given by the concrete

being exceeded, the response is stiff and defined by the pre-cracking stiffness, GK. After

cracking occurs, the torsional stiffness is significantly reduced. As is shown in Figure 4-3, the

effect of cyclic loading has a large effect on the torsional stiffness whereby the concrete

torsional strength contribution diminishes, and the ultimate torsional strength becomes Ts, the

torsional capacity of the reinforcement. To calculate the torsional twists corresponding to the

cracking and ultimate torsion moments, relationships from Mitchell and Collins (1991) and

Park and Paulay (1975) for the torsional stiffness were used. For the initial stiffness, the

section is represented by an equivalent thin walled tube or hollow section with the same

overall dimensions and is a function of the section geometry. The post-cracking torsion

stiffness is governed by the deformation of the reinforcing. The pre-cracking stiffness was

calculated using:

4 A02 t
GK G (4.3)
po

where GK = the pre-cracking torsional stiffness; G = the shear modulus; A0 = area enclosed by

the centreline of the thin walled tube with thickness, t; and po = perimeter of the area. The

equivalent wall thickness can be found by using:

Page 4-8
3 Acp
t (4.4)
4 p cp

Where Acp = area enclosed by the outside perimeter of the concrete cross section, pcp =

perimeter of the area Acp.

To find the post-cracking stiffness, GKcr, a ratio of the post-cracking stiffness to pre-

cracking stiffness was calculated using methods included in Park and Paulay (1975). This

method uses the aspect ratio of the section, and a modified reinforcement ratio expression.

For the present experiment, the postulated relationship between torsional strength and

torsional twisting incorporates a distinct reduction of the torsional strength due to flexure,

shear and cyclic interaction. The reduced torsion capacity represents the damage that occurred

from Phase 2 transverse loading, whereby high shear forces and bending moments were

present and flexural plastic hinging occurred. The concrete contribution to the overall

torsional strength is rapidly diminished, and the torsion demands are sustained by the

reinforcement. Featured in Figure 4-3(a) is a representation of the torsional demand. Using

the intersection of the torsional demand and idealised torsion-twist capacity values, an

estimation of the expected torsional twists can be found. The following sub-sections describe

the calculation of the torsional strength and stiffness parameters as well as the torsional

demands such that the relationship (Figure 4-3(a)) can be quantitatively examined.

4.2.1 Determination of the torsional demand

The torsion due to eccentrically supported flooring systems on beams can be easily calculated.

However, the added torsional demand that can be caused by a moment resisting reinforced

connection between floors and supporting beams, under relative rotations between the slabs

and beams, is less well understood. Concern has been raised in New Zealand as to the

unacceptable torsion that would be generated by seating hollow-core units using a

compressible backing board and low-friction bearing strip on an external supporting beam
Page 4-9
(OGrady, 2004). This connection detail was experimentally tested in the Lindsay (2004)

experiment, and only minor effects of torsion were observed. The degree of twist did not raise

any issues over the performance of the floor. Elliot et al (1998) discussed the effects of

eccentrically loaded external beams for both the isolation type floor-to-beam connection

(compressible backing board) and a reinforced floor-to-beam connection, such as the

connection presented in this research. Elliot et al (1998) states:

If the beam is tied into the floor slab through the normal arrangement of continuity tie bars
concreted into some of the opened cores in the floor slab, the torsional stress is virtually
eliminated. There are two reasons for this: (1) a reaction force R generated in the floor plate
will prevent the top of the beam from experiencing inward deflections and (2) the eccentricity
of the load is reduced because of an extended bearing at the end of the cast-in-place infill.

This statement is fundamentally flawed when lateral displacement conditions are considered.

The relative rotations between the floor and supporting beam induce moments in the floor to

beam connection and these moments generate significant torsion demands at the ends of the

beams. The assumption that the floor remains rigidly connected to the supporting beam, and

rotates with the beam is false. Previous experiments (Matthews, 2004 and Lindsay, 2004) as

well as the current experiment, have shown that under lateral displacements when the

supporting beam rotates, the floor slab stays predominantly horizontal. This relative rotation

between the floor and supporting beam induces a force couple within the reinforcement

connecting the floor and beams. Figure 4-4 shows the falsely claimed behaviour and the

observed behaviour of the reinforced beam-floor connections. The induced moments caused

by the relative rotations impose torsion on the beam, as shown in Figure 4-3(b) and Figure

4-4(b). Therefore, the statement made by Elliot et al (1998) that torsionally stresses are

virtually eliminated by using a reinforced composite beam-to-floor connection is incorrect.

Page 4-10
(a) Erroneous claimed behaviour of the (b) Observed behaviour of the reinforced
reinforced seated floor slab connection seated floor slab connection
(Elliot et al 1998)
Figure 4-4 The torsional demand from reinforced hollow-core to floor slab connections

In addition to the torsion moment generated by the reinforced floor-beam connection, the

effects of beam elongation also cause torsional demands. The growth of the perimeter frame

beams, parallel to the floor span, exert tension forces into the flooring, which acts to pull the

floor slabs off the seating of the supporting beams. This means that the reaction force, R,

generated in the floor plate is not present, and the tension force from beam elongation further

adds to the torsional demand.

The torsional demands calculated for this experiment were derived from the results

from two-dimensional sub-assemblage testing (discussed in Section 2). The moment capacity

of the floor-to-beam connection (for one hollow-core unit) was observed from the

experimental results and good comparisons were found between the calculated positive and

negative moment capacity values. This moment capacity was used to find the torsional

moment demand imposed on the transverse beams, as is shown in Figure 4-4(b). Accounting

for the moment generated by the connection between the supporting beam and four hollow-

core units, an applied torsional moment at the beam ends of 79kNm was calculated (same for

both positive and negative moments). For the compressible backing board connection

(Lindsay, 2004), the eccentric loading of the hollow-core flooring on the concrete seat of the

supporting beams does result in torsion under static conditions. Under lateral loading, the

moment induced by the section under relative rotations was minimal as it was observed that

the compressible backing board did not actually compress and a force couple could not be
Page 4-11
achieved. A torsional moment demand from the eccentric floor loading of 20kNm at each end

of the transverse beams was calculated, although it must be noted that this is a lower bound

value as live loading was not included in the testing.

4.2.2 Transverse beam torsional capacity

For Phase 1 and 3, the amount of bending moment and shear force acting at the ends of the

transverse beams would have been small, as determined from the static loading of the floor

slabs. Therefore, the following calculations of the torsional capacity of the concrete transverse

beam do not include flexure and shear actions, as they are sufficiently small to negate. Under

different circumstances, for example a 45 degree earthquake whereby the longitudinal and

transverse directions are simultaneously loaded, then the torsion would have to be considered

together with flexure and shear.

Using the equations and methods presented, the torsional strength and stiffness values

for the transverse beams were calculated. The cracking torsion moment was calculated to be

77kNm and the torsion contribution from the reinforcement was found to be 200kNm. Using

the stiffness formulation presented in Equations 4.3 and 4.4, the pre-cracking stiffness was

95500kNm2 and from methods presented in Park and Paulay (1975), the post-cracking

stiffness was found to be 4% of the pre-cracked stiffness (3800kNm2). Using these calculated

torsional strength and stiffness values, the torsion-twist relationship was generated

(Figure 4-5). This offers a quantitative form of the idealised torsion-twist relationship shown

in Figure 4-3(a).

The representation of the torsional demand and capacity shown in Figure 4-5, can be

used to explain the amount of torsional twisting that occurred within the plastic hinge regions

of the west transverse beams (Figure 4-2(b)). The intersection between the demand for the

current experiment and capacity curves for Phase 1 indicates that there would be a torsional

Page 4-12
rotation of around 0.0064 radians (0.64%). This is approximately true for the experimental

discrepancies seen during the later stages of Phase 1, where at the first and second cycles to

1%, the difference between the rotation of the column and the supporting beam adjacent to

the plastic hinges was around 0.7%. At the first and second cycle peaks to -2% drift, the

torsional twist at the ends of the west transverse beam was around 1%. The reduced stiffness

for Phase 3 longitudinal re-loading indicates that significantly higher torsional twisting

occurred. The intersection point between the capacity and demand for Phase 3 shows that

torsional drifts of around 0.021 radians (2.1%) could be expected. During the later stages of

the experiment, the ends of the beams were experiencing rotations of over 4%. The calculated

value of 2.1% underestimates the amount of torsional twisting of the torsional hinges.

However, the suggested representation still provides a relatively good approximation of the

large rotations that can occur due to torsion.

300

250
Torsion Moment (kNm)

200

150

100 Phase I
Phase III
50 Current Experiment Demand
Lindsay (2004) Experiment Demand
0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
Torsional Twist (rads/m)

Figure 4-5 Representation of the calculated torsional strengths and demands

The difficulty in accurately assessing torsion quantitatively means that it is difficult to

confidently predict torsional performance. Uncertainties arise due to the difference between

various methods in calculating the torsional capacity of section as well as the assessment of

the torsional demand on the supporting beams. However, the calculations provided a

reasonably accurate insight into the amount of torsional rotations of the supporting beam

Page 4-13
plastic hinge zones that can occur. However, the fact that the simple calculations produce

estimates of torsional rotation that are less than what was shown by the actual experimental

results, means that torsion of floor supporting beams needs to be considered seriously.

4.3 DISPLACEMENT INCOMPATIBILITY

4.3.1 Identified Issues and Previous Findings

The occurrence of vertical displacement differences between flooring systems and lateral load

resisting frame systems has been accepted as a source of damage for floor diaphragms.

However, little analytical and experimental work has been done to assess the incompatibility.

The displacement incompatibility arises due to the different deformation patterns of the

flooring and frame elements, and is schematically shown in Figure 4-6. Under lateral loading,

the frame beams deform in double curvature, whereas flooring systems, that span parallel to

the frame, are most commonly designed to act as though simply supported. In cases where the

flooring is adequately tied into the beams and is forced to deform in an undesirable double

curvature pattern, the floor slab is prone to damage and effective load paths and transfer of

inertia forces are jeopardised.

Displacement Displacement
incompatibility incompatibility
between the frame between the frame
and h/c unit and h/c unit

Hollow-core unit
sags and deforms Frame deforms in
in single curvature double curvature

Figure 4-6 Illustration of the causes of displacement incompatibility (Matthews, 2004)

In previous experiments, damage due to displacement incompatibility has been reported.

Laus (2002) experiments observed damage between the perimeter beams and floor and

concluded that the vertical movement of the floor relative to the beams has the potential to

cause failure of the slab. Displacement incompatibility was evident due to damage that

Page 4-14
occurred during the hollow-core super-assemblage experiment conducted by Matthews

(2004). A strong connection between the first hollow-core unit and adjoining perimeter beams

(Figure 4-7(a)) forced the floor to deform in double curvature and these deformations could

not be sustained by the webs of the hollow-core unit adjacent to the perimeter beams.

Ultimately the localised over-stressing resulted in the collapse of the bottom half of the first

hollow-core unit beside the main frame. The adverse deformations also resulted in fracturing

of the cold drawn mesh reinforcing the topping slab and caused an unzipping tear

longitudinally between the first and second floor units which can be seen in Figure 4-7(c).

This tear was first noticed when cracking was observed adjacent to the centre column at

0.75% inter-storey drift, and by 1.95% all of the cold drawn mesh crossing this crack had

fractured.

Zone of Grade 430 starters Zone of


weakness at 300 crs & 665 mesh 75mm topping weakness
in 75mm topping slab

750mm
First timber
Second Hollow-core First Hollow-core Unit
infill
Hollow-core Unit
Unit Perimeter beam
Perimeter beam
Column face
Column face

(a) Lateral connection used in the Matthews (b) Lateral connection used in the Lindsay
(2004) experiment (2004) experiment

(c) Tearing damage observed during (d) Infill floor damage witnessed during the
Matthews (2004) testing Lindsay (2004) experiment
Figure 4-7 Lateral perimeter beam floor unit connection details and damage.
Page 4-15
The Lindsay (2004) experiment investigated future construction details addressing the

problems observed during the Matthews (2004) test. The lateral connection between the

perimeter beams and initial hollow-core floor unit consisted of a topping slab cast over a

timber plank infill 750mm wide, as shown in Figure 4-7(b). This was intended to be a flexible

link isolating the two structural systems to accommodate the vertical displacement

incompatibilities while still being able to transfer inertia forces from the floor to the frame.

Figure 4-7(d) shows the damage incurred by the infill slab during the testing. The solution

was able to satisfactorily withstand the displacement incompatibility and damage to the

flooring was minimised. However the premature termination of the starter bars at the interface

between the infill and first hollow-core unit created a zone of weakness where tearing was

observed and the ductile mesh reinforcing the topping slab fractured.

The testing of precast seismic structural systems in the PRESSS research program (Priestley

et al 1999) included connection solutions designed to overcome the problems associated with

displacement incompatibility. The welded X plate connectors, such as those shown in

Figure 4-8, connected the double tee flooring to the perimeter frame at discrete points on the

lower levels of the structure. The X plate connectors were designed by capacity design

principals to act as an energy dissipating device, and limit the amount of inertia forces

transferred to the perimeter moment resisting frame.

Figure 4-8 Welded X-plate connecting device

Page 4-16
4.3.2 Analytical modelling of displacement incompatibility

The analysis involved the construction of an analytical model in the inelastic time history

program Ruaumoko (Carr, 2002), utilising a plasticity concentrated approach. A two-

dimensional model was set up to model the longitudinal lateral loading of the front frame, and

consisted of members representing the frame and floor elements and springs connecting the

floor to the frame. Different spring members were used to model the level of fixity of the end

support conditions, and along the frame between the frame and flooring. The model was

calibrated by assessing the base shear capacity of the system and comparing it to the

experimental values.

Figure 4-9 shows the displacement incompatibility between the longitudinal perimeter

beams and the adjacent parallel hollow-core unit from the experiment and modelling for both

the Lindsay (2004) and current experiments. The lateral spring properties modelling the infill

connection were kept the same, but the end spring properties were altered to model the pinned

type seating connection for Lindsay (2004) and the rigid connection in the present

experiment. The results show promising correlation between the experiment and analytical

model. The overall deformation trends are matched although the model slightly

underestimates the amount of incompatibility that occurs.

After gaining satisfactory comparisons with the experiment, changes to various

parameters of the model allowed other areas of interest to be investigated. Sensitivity analyses

were performed examining the effect of: different end support conditions, different lateral

connection details, different bay lengths and numbers of bays and different flooring types.

The effects of the different parameters was seen by examining both the displacement

incompatibility patterns as well as the overall system base shear capacities, and the results

from these analyses are presented in Appendix D.

Page 4-17
12
Displacement Incompatibility 10
8
6
4 Position along fram e (m m )
2
(mm)

0
-2 0 2000 4000 6000 8000 10000 12000

-4
-6 Experiment 1%
Model 1%
-8 Experiment 2%
-10 Model 2%

-12

(a) Observed and modelled displacement incompatibility from Lindsay (2004)


12
10
Displacement Incompatibility

8
6
4
2 Position along fram e (m m )
(mm)

0
-2 0 2000 4000 6000 8000 10000 12000

-4
-6 Experiment 1%
Model 1%
-8 Experiment 2%
-10 Model 2%

-12
(b) Observed and modelled displacement incompatibility from the current experiment
Figure 4-9 Observed and modelled displacement incompatibility

From the sensitivity analyses, it was seen that the end connection detail, through the stiffness

of the end spring, had a large influence on system strength capacity and displacement

incompatibility values. The base shear capacity for the system was 25% higher for the rigid

connection than the pinned and semi-pinned connections. This is matched to a certain extent

by the higher ultimate base shear capacity of the current experimental structure than the

equivalent Lindsay (2004) base shear capacity. For the rigid end conditions, the maximum

displacement incompatibility occurred on both sides adjacent to the central column, as the

floor unit deformed in single curvature due to the fixity at the ends. For the pinned end

Page 4-18
conditions, which allowed the floor to simply sag, displacement incompatibility peaks were

seen close to the supports as well as either side of the central column, and the displacement

incompatibility values were generally higher than for the fixed end case.

The computational modelling conducted has more recently been extended by work

completed at the University of Canterbury through several research projects. Taylor (2004)

used results from sensitivity analyses of the controlling parameters, to find the points of

minimum displacement incompatibility. For each of the influencing parameters, a list of the

non-dimensional lengths at which displacement incompatibilities are a minimum have been

made to be able to form the basis for the location of connectors. This work preceded work by

McKenzie (2004), who investigated the importance of higher mode inertia forces and if

energy dissipating discrete connectors were able to reduce the demand imposed on lateral load

resisting systems. Jensen (2004) analysed the diaphragm action due to the effects of different

locations and numbers of floor-frame connectors. More favourable conditions were seen for

several connectors as the inertial forces could be distributed over a wider area, however the

location of the connectors is an important parameter considering the displacement

incompatibility issues discussed.

4.3.3 Scope for future advancements

A recent change in direction in earthquake engineering structural research into the

development of post-tensioned rocking frame and wall solutions involved with damage

avoidance design principles has required accompanying attention to floor systems. In

particular, research has focused into isolating the floor and lateral load resisting elements to

limit damage, while still maintaining a connection capable of transferring inertia and

structural interaction forces from the floor to the lateral load resisting system. Work has been

done testing various mechanical connectors and devices that can join floor and frame

Page 4-19
members, and one such connector is shown in Figure 4-8. Connectors that possess energy

dissipation qualities, which are able to limit the inertia forces being transferred to a perimeter

lateral load resisting frame, which in turn restricts damage, have previously been developed

and tested. Naito et al (2004) modelled connectors for double tee floors and found that

connectors designed according to capacity design principals performed in a ductile and

desirable manner. The welded X plate type connectors used in the PRESSS program, as

shown in Figure 4-8, use shear forces to produce a flexural response, which increases the

ductility and significantly improves performance (Priestley et al, 1999).

Despite this previous research, similar work investigating the viability of systems

involving these connectors has been minimal in New Zealand and research should look to

develop systems tailored for New Zealand construction practice. The timber infill solution has

proven to be a successful way to overcome the displacement incompatibility issue. However,

in other building situations, the timber infill may be undesirable, and a flat-soffit option with

flooring butted up against the perimeter frame may be required. In these cases, discrete

connectors that allow vertical displacement but are fixed horizontally transferring inertia

and structural interaction forces and dissipating energy, would be required.

4.4 LATERAL LOAD CAPACITY OF THE SYSTEM

This section discusses the calculation of the lateral load capacity of the super-assemblage and

the comparisons of the predicted strengths with the actual behaviour. The lateral load capacity

of the system was found by incorporating all sources of strength enhancement, including the

starter bar and beam-to-floor slab reinforcement, diaphragm reinforcement and the influence

of the floor slab through the bowstring effect. Full calculations of the lateral load capacity

values are provided in Appendix D.

Page 4-20
4.4.1 Bowstring Effect

The bowstring effect occurs as a three-dimensional consequence of beam elongation. The

floor system provides restraint due to the growth of the concrete frames. This restraint causes

large compression forces in the frame beams, which significantly increases the moment

capacity of the frame. Tensile forces are induced in the flooring, which are transferred to the

frame beams in a compression arching manner, which explains the name the bowstring

effect, where the tension in the floor acts as the string, and the compression arch represents

the bow. Fenwick et al (1999) demonstrated that the presence of floor slabs parallel to

perimeter frame beams can cause a significant increase in the negative moment capacity of

the beams. The bowstring effect is further explained by Bull (2003) and was observed during

experiments conducted by Matthews (2004) and Lindsay (2004). The bowstring effect was

also graphically evident during the present experiment. Cracking within the topping slab over

the timber infill and hollow-core floor units near the east and west ends of the structure,

arched inwards towards the south centre column. The crack distribution followed the same

compression arching pattern of the bowstring effect as shown by Figure 4-10.

centroid of compression arch


compression struts

end
connection
eccentricity (e) reinforcement

2e

centroid of tension tie


uniformly distributed load due to
diaphragm reinforcement and drag
bars

Figure 4-10 Bowstring effect present in the floor slab

Page 4-21
Figure 4-10 shows the location of the compression arch and the centroid of the tension tie in

the floor. The depth of the compression arch, the eccentricity, is the distance between the

centroid of the tension tie in the floor, and the centroid of the compression force at the inside

edge of the south perimeter beams. The tie force is a combination of the capacity of the

reinforcement at the hollow-core to support beam connection as well as friction of the floor.

This resultant tie force acts at a distance of the eccentricity and is a combination of the

attributing tension forces from the end connection reinforcement over a width of two times

the eccentricity. The tie force can be found from the following:

2eAs f y
T  2PWle (4.5)
s

where e = midspan eccentricity of the tie force with respect to the centroid of the compression

strut (2e is the width that contributes to the combined tie force); T = combined tie force;

l = overall span of arch (the length of the structure for this case); As = end connection

reinforcement area (starter bars and hairpin bars); fy = end connection reinforcement yield

strength; s = end connection reinforcement spacing; P = coefficient of friction for the bearing

strip; and W = weight of the hollow-core.

By equilibrium, the tension force in the floor slab is transferred to the perimeter beams

by diagonal compression struts. These diagonal struts follow the observed crack pattern in the

arching pattern demonstrated in Figure 4-10. The eccentricity can be found by considering the

diaphragm reinforcement orthogonal to the tie force. The moment generated by the force

couple between the floor slab and perimeter beams can be equated with the moment generated

by the distributed load given by diaphragm reinforcement orthogonal to the tie force, w. The

compression arch is equivalent to the moment from the uniformly distributed load given by

the diaphragm reinforcement, which is in the shape of a parabola (as is the case for simply

supported beams under distributed loads). The mechanism can be expressed as follows when

considering equilibrium:

Page 4-22
wl 2 Pl
Te  (4.6)
8 4

Where the first and second parts of (the right hand side of) Equation 4.6 respectively represent

the uniform distribution of slab topping reinforcement along the entire floor span, and the

drag bars anchored into the central column. Equation 4.6 can be expanded to give:

Asd f yd l 2 Ast f yt l
Te  (4.7)
8s sd 4

where Asd = area of one bar of the diaphragm reinforcement; fyd = yield strength of the

diaphragm reinforcement; ssd = spacing of the diaphragm reinforcement; and Ast = area of one

drag bar; fyt = yield strength of the drag bar reinforcement

This formulation provides an admissible mechanism from which the eccentricity can

be found (by combining Equations 4.5 and 4.6), and therefore the combined tie force and by

equilibrium, the compression axial force induced in the south perimeter beams can be

calculated. This solution provides an upper bound method as it assumes that the reinforcement

at the end supports of the hollow-core floor slabs, and diaphragm reinforcement orthogonal to

the floor span are yielding.

Using these equations, and the measured material properties of the reinforcement, an

eccentricity of 2.5m was found. From strain gauge information of the drag bars tying the

interior column into the floor plate, it was observed that the reinforcement was yielding, and

the inclusion for this reinforcement in the mechanism analysis (Equation 4.6) was verified

(see Appendix D for strain information for the drag bar reinforcement). The calculated value

of the depth of the compression arch or eccentricity lies over the middle of the second hollow-

core unit. The location of the calculated tie force and depth of the compression arch matches

well with the cracking pattern witnessed throughout testing as is shown by 4-11(a) whereby a

schematic of the bowstring effect is superimposed on a photo of the floor. During testing, a

longitudinal crack over 7m long formed at the interface between the first and second hollow-

Page 4-23
core units, close to the calculated centroid of the tie force, which indicates that the calculated

eccentricity for the mechanism is supported by the observed behaviour. Using the calculated

eccentricity value, a combined tension force distributed over 5m, of 1254kN was found. By

equilibrium, this tension force is resisted by a compression force of the same magnitude,

which, in this case is experienced by the central column (at the top of the arch). Figure

4-11(b) shows the internal and external actions for the plastic hinges adjoining the south

centre column. The compression force, N, which is transferred through the topping slab into

the perimeter beams, has a significant effect the strength of the plastic hinges of the southeast

and southwest beams framing into the central column.

Compression Arching between the Compression Arching following


floor diaphragm and perimeter the observed crack pattern
beams

N N
Eccentricity = e
2 * Eccentricity = 2e

T
Centroid of the tension tie in the floor
diaphragm

(a) Observed bowstring effect


South
Centre
Column
Axial force from
the bowstring effect,
acting at the level of
the topping slab
N T N
C S +CC
D (d-d') M M (d-d')
C S +CC
T
(d')
South West Beam South East Beam
-
M = (A sfy (d-d'))+(N(d)) +
M = A sfy (d-d')

(b) Moment strength enhancement due to the bowstring effect


Figure 4-11 The influence of the bowstring effect

Page 4-24
As it is displayed in Figure 4-11(b) the negative moment capacity of the plastic hinges

framing into the south centre column are increased by the axial compression force which is

shown to act at the level of the topping slab. The moment contribution from both of the plastic

hinges framing into the interior column can be found by the following:

M beams ( As  As ) f y (d  d ' )  Nd (4.8)

In which As+ = area of tension reinforcement; As- = area of compression reinforcement; fy =

yield strength of tension reinforcement; d = the distance from the extreme compression fibre

to the centroid of the tension reinforcement; d = the distance from the extreme compression

fibre to the centre of the compression block; and N = the axial compression force due to the

bowstring effect.

Using the moment capacity Equation 4.8, the total moment from both of the southeast

and southwest beams hinges (joining the south centre column) experienced by the column is

1859kNm. This is compared to 1004kNm when the bowstring effect in the floor slabs is not

included.

The compression axial forces due to the bowstring effect in this case have been

represented at the height of the topping slab although it is acknowledged that the distribution

of the compression axial forces is more complex than what has been described. The inclusion

of the bowstring effect has been done by way of a simplistic mechanism approach, although it

is acknowledged that it is an upper bound method. Due to this, a significant increase of the

strengths of the perimeter beam plastic hinges was noted due to the compression axial force.

The following sub-sections include the bowstring effect into calculations of the total super-

assemblage lateral load capacity and compare these predictions with the experimental results.

Page 4-25
4.4.2 General System Strength

An important part of finding the capacity of the structure is assessing the negative moment

capacity enhancement provided by tension reinforcement within the topping slab. Matthews

(2004) provided a method of calculating the activated slab width and accompanying strength

from the starter bar reinforcement. This method was based on the experimental findings from

the Matthews (2004) experiment and is a function of the specific structure. It provided an

accurate means of calculating the lateral load capacity of the system. Comparisons were made

with the equivalent methods prescribed in New Zealand Concrete Structures Standard (NZS

3101:1995) and the American Concrete Code (ACI 318-02) and it was found that these

methods significantly underestimated the amount of strength contribution from the activated

tension flange. Undervaluing the amount of moment capacity enhancement creates the

potential of an undesirable strong beam-weak column mechanism forming, and it is therefore

important that the slab and diaphragm reinforcement is accounted for correctly.

The method used by Matthews (2004) characterises the progressive activation of

starter bars according to increasing lateral displacement. Two distinct points are required to

define the width of slab activation: the onset of yielding, and full plasticity. These points were

found by examining the crack widths crossing the interface where starter bar or diaphragm

reinforcement is present. The torsional stiffness of the beam is important in determining the

rate of activation of the starter bar reinforcement. For beams of relatively low torsional

stiffness, the starter bars in the middle of the beam will be activated much later from those at

the end. In this experiment, it was observed that the beams formed torsional hinges at the

ends, and can be classified as having low torsional stiffness. However, a sizeable crack

formed in the topping concrete at the beam-to-floor interface and it was determined that the

initial yield point and point of full plastification were similar to that assumed by Matthews

(2004).

Page 4-26
For this experiment, the slab width at the first yield point was found to be the width of the

timber infill (0.75m), which incorporates three starter bars or longitudinal diaphragm bars,

and full plasticity was reached at 2% drift with an activated slab width of 3.05m,

approximately half the width of the bay. This effective slab width is the same as what was

used by Matthews (2004) and Lindsay (2004) and Figure 4-12 shows the contributing slab

widths.
Activated tension flange Activated tension flange

Beam + starters Beam + starters

JD

Centre of the Centre of the


compression force compression force
(a) First Yield (b) Full Plastification

Figure 4-12 Activated slab width for negative moment enhancement.

The overall strength of the system observed from the experiment is shown in

Figure 4-13 for both the longitudinal and transverse loading phases. The theoretical pushover

curves are superimposed to assess the validity of the theoretical calculations. This predicted

strength capacity incorporates the nominal and overstrength strengths of the beams, starter bar

contributions during the progression from yielding to full plasticity, the effect of the end

support beam-to-floor slab connection reinforcement and the axial compression forces

induced by the bowstring effect. As it can be seen by examining Figure 4-13(a) for both Phase

1 and 3 loading, the correlation is satisfactory. The overstrength capacity strengths (assumed

as 125% and 140% of the nominal strengths) have been plotted as upper bound values and

actual strength of the system lies between the two overstrength predictions. It can be seen that

the later cycles are well predicted; the 3% and 4% cycle peaks or maximum strengths are

slightly underestimated by the nominal theoretical values but well matched by the 40%

overstrength prediction.

Page 4-27
The base shear capacity for the transverse loading phase is shown by Figure 4-13(b).

The accompanying theoretical pushover curve shows good agreement with the experimental

results. For positive drifts, the experimental results experimental values fall in between the

upper bound overstrength capacity values and the lower bound nominal strengths. The

negative drift predictions conversely, are higher than the actual results. This could signify that

there was less starter bar reinforcement activated than what was assumed during calculating

the pushover curve. The experimental values graphed are for before and after a torsion test

that was conducted mid-way through Phase 2 loading. It is discussed further in Appendix D,

but it can be seen that the torsion test undertaken did not cause a reduction in the load

carrying capacity of the system.

It should also be noted that a yield inter-storey drift of 0.64% was used for the yield

strengths. As the hysteretic behaviour indicates, this theoretical yield drift compares well with

the observed experimental yielding point and elastic stiffness. A more in depth assessment of

the super-assemblage stiffness as well as details of the calculation of the yield drift are located

in Appendix D.

4.4.2 Localised Strengths

The individual or combined plastic hinge strengths are shown by the moment rotation

relationships in Figure 4-14. The plastic hinges of the south frame for longitudinal loading are

shown in Figure 4-13(a) for both the southeast and southwest hinges, and the combination of

the centre hinges framing into the centre column. The general agreement is good with the later

cycle peaks lying between the nominal and overstrength estimates. The comparison between

the observed experimental behaviour theoretical predictions for the central column is good,

which indicates that the method, albeit an upper bound approach, used to quantify the added

moment due to the bowstring effect is sufficiently accurate.

Page 4-28
1800
1600

Base Shear
Force (kN)
1400
1200
1000
800
600
400
200
Inter-Storey Drift (%)
0
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
-200
-400
-600
-800
Experiment Phase 3
-1000
Experiment Phase 1
-1200
Nominal Prediction
-1400 25% Overstrength Prediction
-1600 40% Overstrength Prediction
-1800

(a) Phase 1 and 3 Longitudinal loading: Experimental base shear capacity and theoretical
pushover curve.
1200
Base Shear
Force (kN)

900

600

300

Inter-Storey Drift (%)


0
-4 -3 -2 -1 0 1 2 3 4

-300

-600 Experiment (pre-torsion)

Experiment (post-torsion)

-900 Nominal Prediction

25% Overstrength Prediction

-1200

(b) Phase 2 Transverse loading: Experimental base shear capacity and theoretical pushover
curve.
Figure 4-13 Base shear capacity: experimental and theoretical relationships.

Page 4-29
1200 3000 1200

Moment (kNm)
Moment (kNm)

Moment (kNm)
1000 2500 1000

800 2000 800

600 1500 600

400 1000 400

200 500 200


R o t a t io n ( %) R o t a t io n ( %) R o t a t io n ( %)
0 0 0
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 -6 -5 -4 -3 -2 -1 0
-500 1 2 3 4 5 6 -6 -5 -4 -3 -2 -1 0
-200 1 2 3 4 5 6
-200
-400 -1000 -400

-1500 Phase 1 -600 Phase 1


-600 Phase 1
Phase 3 Phase 3
Phase 3 -2000 -800
-800 Nominal Strengt h Nominal Strengt h
Nominal Strength 25%Overst rengt h
-1000 -2500 -1000 25%Overstrengt h
25%Overstrength
-3000 -1200
-1200

Southwest hinge, south beams South centre hinges Southeast hinge, south beams
(a) Phase 1 Longitudinal loading: plastic hinge strengths
1200 1200
1000
Moment

1000

Moment
(kNm)

(kNm)
800 800
600 600
400 400
200 200
Rotation (%) Rotation (%)
0 0
-4 -3 -2 -1 -200 0 1 2 3 4 -4 -3 -2 -1 -200 0 1 2 3 4

-400 -400
Pre-Torsion Pre-Torsion
-600 -600
Post-Torsion Post-Torsion
-800 Nominal Strength -800
Nominal Strength
-1000 25% Overstrength -1000 25% Overstrength
-1200 -1200

North hinge, West transverse beam North hinge, East transverse beam

1200 1200
1000
Moment

1000
Moment

(kNm)
(kNm)

800 800
600 600
400 400
200 200
Rotation (%) 0
Rotation (%) 0
-4 -3 -2 -1 -200 0 1 2 3 4 -4 -3 -2 -1 -200 0 1 2 3 4

-400 -400
Pre-Torsion Pre-Torsion
-600 -600 Post-Torsion
Post-Torsion
-800 -800 Nominal Strength
Nominal Strength
-1000 -1000 25% Overstrength
25% Overstrength
-1200 -1200

South hinge, West transverse beam South hinge, East transverse beam
(b) Phase 2 Transverse loading: Individual hinge strengths.

Figure 4-14 Individual hinge strengths: experimental and theoretical pushover relationships.

Experimental and theoretical comparisons for the transverse loading are provided in Figure

4-14(b). It is shown that the theoretical strength model represents the actual behaviour well.

The positive moment capacities for each of the hinges are matched well by the overstrength

moment predictions. However, there is a slight over-prediction of the negative moment

Page 4-30
capacities for the southeast and southwest hinges of the transverse beams. The nominal

moment capacity values provide a more suitable match in this case. This overestimation could

mean that the amount of activated slab contributing is less than what was specified; half of the

longitudinal south bays at full plasticity.

4.5 BEAM ELONGATION

One of the significant outcomes from the Matthews (2004) experiment was the development

of a rational mechanics based theory to predict beam elongation. A rainflow counting method

was been employed to account for the growth of plastic hinges in a frame system under

inelastic rotations. The theory is briefly explained in Section 2 of this thesis, but further

information is provided by Matthews (2004). The calculations made for the theoretical

predictions of beam elongation are tabulated in Appendix D.

Figure 4-15(a) and (b) displays the observed and theoretical beam elongation results

for Phase 1 and 3 longitudinal loading. The elongation values represented are for the total

frame and a consideration of the pull-off demands imposed on the floor seating connections

require the values shown to be halved. As it is illustrated, the observed and predicted results

agree well, especially during Phase 1. The theoretical elongations slightly underestimate the

actual values up until the first cycle to 4% during Phase 3, after which the experimental

values level off and even decrease. The elongation predictions of the final loading cycles to

5% are not reached due to instrumental errors, but despite this, the agreement is satisfactory

up until the 4% cycles. After Phase 1 loading, the permanent elongation was 32mm, from

which the elongation continued to grow during Phase 3 re-loading and at the end of testing,

the permanent elongation was 72mm.

Page 4-31
Figure 4-15(c) shows the frame beam elongation results obtained from the Phase 2

transverse loading. The predicted values display elongation for the outside transverse frames

and are displayed by Figure 4-14(d). The beam elongation theory once again predicts the

actual behaviour well. The actual results are marginally lower than the predicted values,

however the difference is small. At +4% the predicted beam elongation was 30mm and the

recorded experiment elongations were 29mm and 30mm for the west and east bays

respectively.

120 120
Elongation

Elongation
T heor y
(mm)

(mm)
100 100

80 80

60 60

40 40

Phase 1 20
20
Phase 3
0 0

-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
Inter-storey Drift (%) Inter-storey Drift (%)

(a) Phase 1 and 3 longitudinal experimental (b) Phase 1 and 3 longitudinal theoretical beam
beam elongation elongation
35 35
Elongation
Elongation

T heor y
(mm)
(mm)

30 30

25 25

20 20

15 15

10 10
East Bay
5 5
West Bay
0 0
-4 -3 -2 -1 0 1 2 3 4 -4 -3 -2 -1 0 1 2 3 4
Inter-storey Drift (%) Inter-storey Drift (%)

(c) Phase 2 transverse experimental beam (d) Phase 2 transverse theoretical beam
elongation elongation
Figure 4-15 Experimental and theoretical beam elongation results.

Page 4-32
4.6 LOW CYCLE FATIGUE OF THE REINFORCEMENT

The fracture of the reinforcement observed during the later stages of the experiment can be

predicted by low cycle fatigue theory. The fracture of the longitudinal reinforcement during

the Lindsay (2004) experiment was able to be predicted to a satisfactory level of accuracy.

Low cycle fatigue is also addressed in Section 2 for the sub-assemblage tests. The low cycle

fatigue theory enables the plastic rotation life capacity of the reinforcement to be calculated

and is expressed as follows (Dutta and Mander, 2001):

D L p 1 L p
2 N f
Tp 0.16 1 
 0.5
(4.9)
D' D 2 L

where Tp = plastic drift; L = lever arm of the cantilever column; D = overall member depth;

D = distance between outer layers of steel; 2Nf = the number of reversals to the appearance of

the first fatigue crack with which Nf is the effective number of constant amplitude cycles for a

variable amplitude displacement history; and Lp = equivalent plastic hinge length, which is

given by:

Lp 0.08 L  4400H y d b (4.10)

in which Hy = the yield strain and db = the diameter of longitudinal reinforcement.

Miners well-known linear damage accumulation theory can be utilised to calculate

the equivalent number of equi-amplitude cycles (Nf):

2
T
Nf i T

(4.11)
ref

where T = drift for the ith cycle of loading and Tref = reference drift for an equivalent constant

amplitude loading history. Using a reference drift of 5% an applied cyclic demand of Nf = 4.5

cycles is calculated for the experiment. This signifies that 4.5 equivalent cycles of 5% drift

caused the low cycle fatigue fracture of the reinforcement. By re-arranging Equation 4.9, the

Page 4-33
theoretical equivalent number of equi-amplitude cycles can be found. Using a reference drift

of 0.05 radians, a yield drift of 0.0064 radians and a equivalent plastic hinge length of

511mm, an equivalent number of 3.6 cycles to 5% was calculated This underestimates the

first fracture, which occurred after 4.5 cycles. Given that conventional wisdom considers that

any result that falls within 25% of the predicted value is satisfactory, it is evident that the

experiment (NF-Expt = 4.5) and theory (NF-Theory = 3.6) match to a satisfactory level.

4.7 COMPARISONS WITH PREVIOUS RESEARCH

The research presented in this thesis report follows work that has identified the seismic

inadequacies of precast hollow-core floor systems and investigated solutions to address the

identified problems. With the results and understanding gained from each of the three super-

assemblage experiments and associated two-dimensional sub-assemblage investigations, the

nuances of hollow-core behaviour have become clearer. There has been a progression of

design solutions and correspondingly better performance. This experiment, as it was aimed to

do, has answered many questions and proven the recommendations given by previous

researchers. This sub-section briefly assesses aspects of this experiment in relation to the

investigations by Matthews (2004) and Lindsay (2004) as a way to conclude this particular

line of hollow-core research.

Figure 4-16 shows the lateral load demand imparted on the structure during each of

the three super-assemblage experiments as well as the timing of significant failure events. As

it can be seen, the cyclic demand for the Lindsay (2004) and current investigations of new

construction details, was more onerous than the original Matthews (2004) experiment, which

examined of existing buildings. The location of failure events also signify improved

performance throughout the three experiments, with the later two experiments being able to

sustain much higher drift amplitudes and number of cycles.

Page 4-34
5 5 5
4 Initial Mesh Partial

Inter-storey Drift (%)

Inter-storey Drift (%)


4 4
Inter-storey Drift (%)

3 Fracture 3 3
Collapse
2 2 2
1 1 1
0 0 0
-1 -1 -1
-2 -2 -2
-3 -3 -3
-4 -4 -4 Total Collapse
-5 -5 -5

(a) Loading protocol used for the Matthews (2004) experiment

5 5 5
Initial Mesh Low cycle fatigue
4 Inter-storey Drift (%) 4 4 Fracture of main beam bars
Inter-storey Drift (%)

Inter-storey Drift (%)


3 3 3
2 2 2
1 1 1
0 0 0
-1 -1 -1
-2 -2 -2
-3 -3 -3
-4 -4 -4
-5 -5 -5

(b) Loading protocol used for the Lindsay (2004) experiment

5 5 5
Low cycle fatigue
4 4 4 of main beam bars

Inter-storey Drift (%)


Inter-storey Drift (%)

Inter-storey Drift (%)

3 3 3
2 2 2
1 1 1
0 0 0
-1 -1 -1
-2 -2 -2
-3 -3 -3
-4 -4
-4
-5 -5
-5

(c) Loading protocol used for the current experiment


Figure 4-16 Loading cycles imposed and observed failures for the current and previous
experiments.

4.7.1 Performance of the Seated Connection

Figure 4-17 shows a schematic representation of the behaviour of the hollow-core-to-support

beam seated connections for each of the three tests. The seated connection used for the

Matthews (2004) experiment was originally assumed to slide on the concrete seating ledge

relative to the supporting beam. However, the rigidity of the mortar bed seating surface did

not allow the floor unit to slide, and under cyclic loading, the relative rotations between the

supporting beam and floor slab resulted in large cracks forming in webs of the hollow-core

unit. This eventually caused a loss of seating failure of the hollow-core flooring. The floor-

beam connection tested by Lindsay (2004), as a result of the previous poor performance, was

Page 4-35
a pinned type connection, which featured the hollow-core flooring seated on a low friction

bearing strip to allow sliding and a compression backing board. This detail behaved well,

although spalling of the concrete ledge concrete and sliding of the low friction bearing strip

relative to the supporting beam occurred late in the experiment.

The beam-to-floor connection used in this current experiment rigidly tied the floor into

the beam, possessing a degree of redundancy not offered by the Lindsay (2004) detail.

Improved performance under relative rotation was observed due to the promotion of a critical

section at the beam-slab interface at which the rotation was concentrated. The low friction

bearing strip performed well and allowed the flooring to slide relative to the supporting beam

under the effects of beam elongation, and after the completion of testing, the hollow-core

floor units had suffered very little damage.

Compressible material
does not compress
Column

Snapping action

Unit slides

Assumed to slide Actual behaviour

(a) Matthews (2004) seated connection (b) Lindsay (2004) seated connection
behaviour. behaviour.
Rotation centred at Hollow-core units
critical section remain undamaged

Moment resisted by connection


reinforcing force couple
Unit sliding on
low friction
bearing strip
(c) Current experiment seated connection behaviour.
Figure 4-17 Observed behaviour of the floor-to-beam under relative rotation.

Page 4-36
The successive experiments have seen dramatic improvements of the seated connection

performance details. The problems associated with the inadequate behaviour of existing

building details has been remedied in terms of future construction and the seismic adequacy

of the two latter supporting connection details that have been added to the New Zealand

Concrete Code Amendment NZS 3101:2004 have been experimentally verified.

4.7.2 Performance of the Perimeter Beam to Hollow-Core Connection

The lateral hollow-core connection throughout the current and previous experiments has been

one of the primary sources of examination and details from the previous tests are shown in

Figure 4-7. A leading source of damage witnessed during the Matthews (2004) test occurred

due to deformation incompatibilities between the perimeter frame and adjoining hollow-core

floor units. The discussion of displacement incompatibility outlines the identified issues and

reasons for the damage. As a solution to these issues, a timber infill solution that allows the

perimeter beams and flooring to displace freely individually was proposed and tested by

Lindsay (2004) and a dramatic improvement was noted. However, the termination of the

starter bar reinforcement prior to the infill-to-floor interface placed high demands on the

ductile mesh reinforcement in the topping slab which fractured at relatively early stages and a

resulting large tear within the topping concrete. To overcome the deficiencies from this

solution, a similar timber infill detail was tested again during the current experiment,

incorporating longer starter bars and conventional reinforcing in place of the ductile mesh

used to reinforce the diaphragm. These enhancements resulted in improved performance

whereby the longer starter bars distributed stresses over a larger area and there was no major

cracking or tearing ensuring that inertia and structural interaction loads from the flooring are

able to be transferred to the perimeter frame.

Page 4-37
4.7.3 Overall performance comparisons

Comparisons between the base shear hysteretic behaviour of the three experiments, shown in

Figure 4-18, are able to show the relative overall strengths and stiffnesses of the systems and

energy dissipation characteristics. However, the overall performance of the structure cannot

be determined from the hystereses relationship as it is dominated by the performance of the

reinforced concrete frame. Upon examining Figure 4-18, the major improvements made by

the current and Lindsay (2004) experiments are clouded by the similar performances of the

perimeter concrete frames. Despite this, it can be seen that similar performance of the

perimeter concrete frames can be achieved by repairing existing concrete frames (Lindsay,

2004) and by achieving equivalent strengths using Grade 500E reinforcing.

In addition to the design changes made following the Matthews (2004) test to the end

and lateral hollow-core connection and the corresponding improved performance, there are

also several other aspects of the test structure behaviour to consider. There was little or no

damage to the southern end corners of the hollow-core units signifying the hooked bar,

utilised in the transverse beams to reduce the risk of relocated plastic hinging from occurring,

performed as desired. Other small modifications made from the Lindsay (2004) such as a

second generation bearing strip and seat concrete transverse reinforcing also proved to be

worthwhile. Interestingly, the pattern of damage within the plastic hinges for the current

experiment was different than the Matthews (2004) and Lindsay (2004) experiments.

Significant cracks, one at the beam-column interface and another 100-200mm away from the

column, formed and progressively opened throughout tests. This contrasts with the previous

experiments and expected behaviour of distributed cracking over a conventional plastic hinge

zone length.

Page 4-38
1600 1200

Base Shear
Force (kN)

Base Shear
Force (kN)
1200 900

800 600

400 300
Inter-Storey Inter-Storey
Drift (%) Drift (%)
0 0
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 -4 -3 -2 -1 0 1 2 3 4
-400 -300

-800 -600

Phase 1
-1200 -900
Phase 3

-1600 -1200

(a) Phase 1 and 3 Longitudional loading (b) Phase 2 Transverse loading


(Matthews, 2004) (Matthews, 2004)
1600 1200
Base Shear
Force (kN)

Base Shear
Force (kN)
1200 900

800 600

400 300
Inter-Storey Inter-Storey
Drift (%) Drift (%)
0 0
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 -4 -3 -2 -1 0 1 2 3 4

-400 -300

-800 -600

Phase 1
-1200 -900
Phase 3

-1600 -1200

(c) Phase 1 and 3 Longitudional loading (d) Phase 2 Transverse loading


(Lindsay, 2004) (Lindsay, 2004)
1600 1200
Base Shear
Force (kN)

Base Shear
Force (kN)

1200 900

800 600

400 300
Inter-Storey Inter-Storey
Drift (%) Drift (%)
0 0
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 -4 -3 -2 -1 0 1 2 3 4

-400 -300

-800 -600

Phase 1
-1200 -900
Phase 3

-1600 -1200

(e) Phase 1 and 3 Longitudional loading (f) Phase 2 Transverse loading


(Current Experiment) (Current Experiment)
Figure 4-18 Comparative hysteretic behaviour for Matthews (2004), Lindsay (2004) and
the current experiment.

Page 4-39
4.8 CONCLUSIONS

The results and discussion provided in this section provide a forensic analysis that enabled an

overview of the super-assemblage performance to be conducted. From the results presented

herein, the following concluding remarks can be made:

1. Due to the added torque created by the rigidly reinforced floor-to-beam connection,

torsional hinging was observed to occur at the ends of the transverse beams. The use of

standard torsional strength and stiffness calculations, to make an assessment of the

torsional performance of the beam sections, was explained. It was seen that the flexural

plastic hinging of the ends of the transverse beams under Phase 2 transverse loading was

largely the reason for the deterioration of the concrete contribution to the torsional

strength, Tc, and hence led to large torsional twists on the softened cracked beam sections

during Phase 3 loading. It is recommended that future research should further address the

complexities of torsional behaviour when combined with concurrent shear and flexural

actions that are expected to occur under realistic bi-lateral seismic repsonse.

2. The construction details altered and improved from the Lindsay (2004) experiment that

features an articulated topping slab cast over a timber infill demonstrated that it

substantial displacement incompatibilities between the hollow-core flooring and

perimeter frame can be sustained. Analytical modelling of the displacement

incompatibility showed good agreement with the experimental results. It was also shown

that extensions of this model could provide a useful tool for analysing and predicting the

performance of different structures as well as possible future solutions.

3. The strength of the super-assemblage was able to be predicted and good correlation was

seen between the theoretical pushover curve and the experimental base shear force

hysteretic behaviour. The method for accounting for the progressive activation of starter

bars with increasing lateral displacement of the structure matched the actual strength

Page 4-40
increase from the initial onset of yielding and the assumed point of full plastification.

The contribution to the lateral load capacity of the front frame due to the restraint

imparted by the floor slab, the bowstring effect, was quantified and seen to be

significant. The observed crack patterns in the floor slab support the concept of a so-

called bowstring effect. This effect induced a substantial axial thrust force in the

perimeter beams which in turn increases the moment demand on the interior column.

4. A summary of the preceding Matthews (2004) and Lindsay (2004) experiments and

current investigation with respect to the comparative overall structural performance was

made. It was seen that there has been distinct improvements through the successive

experiments and larger displacement demands can now be maintained without significant

damage. Several other detailing changes implemented such as conventional reinforcing

in place of mesh, special detailing of the transverse beam, improved bearing strip and

reinforcing of the concrete seat all enhanced the overall performance of the structure.

4.9 REFERENCES

ACI 318-02, Building Code Requirements for Structural Concrete and Commentary,

American Concrete Institute, Farmington Hills, USA.

Bull D.K, 2003, Understanding the Complexities of Designing Diaphragms in Buildings with

Earthquakes, Symposium to Celebrate the Lifetime Contributions of Professors

Emeriti Tom Paulay and Bob Park, Christchurch, pp 93-113.

Carr A.J, 2003, Ruaumoko 2D User Manual, Computer Program Library, Department of Civil

Engineering, University of Canterbury, Christchurch, New Zealand.

Dutta A and Mander J.B, 2001, Energy Based Methodology for Ductile Design of Concrete

Columns, Journal of Structural Engineering, Vol. 127(12), December, pp 1374-1381.

Page 4-41
Elliot K.S, Davies G, Gorgun H, Adlparvar M.R, 1998, The Stability of Precast Concrete

Skeletal Structures, PCI Journal, March-April, pp 42-60.

Fenwick R.C, Davidson B.J, and Lau D, 1999, Strength Enhancement of Beams in Ductile

Seismic Resistant Frames due to Prestressed Components in Floor Slabs, Journal of

the New Zealand Structural Engineering Society (SESOC), Vol 12, No. 1, pp 35-40.

Hsu T.T.C, 1993, Unified Theory of Reinforced Concrete, CRC Press, Boca Raton, USA.

Jensen J, 2004, Diaphragm Action Issues in a Jointed Precast Concrete System, Research

report submitted as part of third professional year coursework, Department of Civil

Engineering, Christchurch, New Zealand.

Lau D.B.N, Fenwick R.C and Davidson B.J, 2002, Seismic performance of R/C perimeter

frames with slabs containing prestressed units, Proceedings for NZSEE conference,

Napier, March, Paper No. 7.2.

Lindsay R.A, 2004, Experiments on the seismic performance of hollow-core floor systems in

precast concrete building, Masters Thesis, Department of Civil Engineering,

University of Canterbury, Christchurch, New Zealand.

Matthews J.G, 2004, Hollow-core floor slab performance following a severe earthquake, PhD

Thesis, Department of Civil Engineering, University of Canterbury, Christchurch,

New Zealand.

Matthews J.G, Mander J.B, Bull D.K, 2004, Prediction of beam elongation in structural

concrete members using a rainflow method, 2004 NZSEE Conference, Rotorua, NZ,

Paper 27, 8 pp.

McKenzie R, 2004, Feasibility Study and Optimum Location of Discrete Floor-Frame

Connectors in a Precast Concrete System, Research report submitted as part of third

professional year coursework, Department of Civil Engineering, Christchurch, New

Zealand.

Page 4-42
Mitchell M.P and Collins D, 1991, Prestressed Concrete Structures, Prentice Hall,

Englewood Cliffs, N.J, USA.

NZS3101, 1995, Concrete Structures Standard, NZS3101, Parts 1 & 2, Standards New

Zealand, Wellington, New Zealand.

NZS3101, 2003, Draft Amendment No. 3 to 1995 Standard (NZS3101), Standards New

Zealand, Wellington, New Zealand.

Naito C.J and Cao L, 2004, Precast Diaphragm Panel Joint Connector Performance, 13th

World Conference on Earthquake Engineering [Proceedings], Vancouver, Canada,

Aug. 2004, Paper 2722.

OGrady C.R, 2004, Letters to the Editor, Journal of the New Zealand Structural Engineering

Society (SESOC), Vol 17, No. 2, pp 8-9.

Park R and Paulay T, 1975, Reinforced Concrete Structures, J. Wiley and Sons Inc, New

York.

Paulay T and Preistley M.J.N, 1992, Seismic Design of Reinforced Concrete and Masonry

Buildings, J. Wiley and Sons Inc, New York.

Priestley M.J, Sritharan S, Conley J.R, Pampanin S, 1999, Preliminary results and

conclusions from the PRESSS five-storey precast concrete test building, PCI Journal,

Vol. 44, No. 6, November-December, pp 42-57.

Taylor L, 2004, Vertical Displacement Incompatibility between Floor Slabs and Seismic

Resisting Systems, Research report submitted as part of third professional year

coursework, Department of Civil Engineering, Christchurch, New Zealand.

Trowsdale J.F, 2004, Seismic Performance of Hollow-core Seating Details Specified by

Amendment No. 3 to NZS 3101:1995, Research report submitted as part of third

professional year coursework, Department of Civil Engineering, Christchurch, New

Zealand.

Page 4-43
5. Section 5 Conclusions and Recommendations

5.1 SUMMARY

The research presented in this thesis continued work done by Matthews (2004) and Lindsay

(2004) into examining the seismic performance of precast concrete structures with hollow-

core flooring. The complete failure of the hollow-core flooring observed during the

investigation by Matthews (2004) showed that existing structures using hollow-core flooring

systems pose a serious risk under seismic events. Continuing research by Lindsay (2004)

showed that with simple detailing enhancements, significant improvement in the seismic

performance of hollow-core floor systems can be expected. The present experimental research

aimed at validating several new detailing enhancements.

Initially, two hollowcore-to-support beam seated connections were tested under

simulated seismic loads at a two-dimensional sub-assemblage level. The support connections

were reinforced and rigidly tied the floor into the supporting beam, the first of which has been

proposed in Amendment No. 3 to the New Zealand Concrete Standard (NZS 3101, 2004). A

newly developed test set-up was described which incorporated both the relative rotation

between the supporting beam and hollow-core floor, and tension in the floor representing the

elongation of a parallel moment resisting frame, as the sources of damage. The sub-

assemblage test specimens were tested up to relative rotations of 5.0% and visual and

instrumental observations from the experiments were outlined.

The main investigation of this thesis involved the testing of a full-scale two bay-by-

one bay concrete frame super-assemblage. The super-assemblage was constructed and several

new detailing improvements recommended from previous research were implemented as well

as the recently proposed, but untested, hollow-core connections with the concrete frame

(NZS 3101, 2004). The full-scale structure was cyclically tested in both the longitudinal and
Page 5-1
transverse directions to inter-storey drifts of 5%. Visual and instrumental observations from

the experiment were presented and discussed.

The third major section of this thesis provided a forensic analytical overview of the

experimental observations. The torsion behaviour of the transverse beams was discussed and

analysed using simple calculations. Displacement incompatibility was also described and

results from the experiment were compared with analytical modelling results. The potential

uses of extending the model were also described. A theoretical pushover curve was generated

predicting the strength of the super-assemblage, and the contribution to the strength of the

moment resisting frame of the floor slab, through the bowstring effect, was quantified.

Finally, as a way of completing the current line of experimental research, the present

investigation was compared and contrasted with previous research conducted by Matthews

(2004) and Lindsay (2004). The comparative performances showed that through improved

understanding of hollow-core performance and accompanying detailing enhancements there

have been significant advancements.

5.2 CONCLUSIONS

From the research presented in this thesis, the following conclusions can be drawn:

1. The three-dimensional testing of a full-scale two bay-by one-bay concrete moment

resisting frame structure incorporating prestressed hollow-core flooring, as designed to

New Zealand Concrete Design Standards (NZS 3101:1995 and amendments) performed

well under simulated seismic loads. The super-assemblage structure was able to sustain

lateral inter-storey drifts up to 5% when the longitudinal reinforcement in the beam

fractured due to low cycle fatigue. The floor system sustained only minor damage, and

life safety was not endangered at any stage during testing.

Page 5-2
2. The reinforced connection between the hollow-core floor slab and supporting beams

performed well up to inter-storey drifts of 5% and at the conclusion of testing, the

damage suffered by the hollow-core floor units was minimal. Desirable action, where the

relative rotations between the floor and supporting beam were concentrated at the beam-

slab interface was achieved. A second generation low friction bearing strip ensured that

the hollow-core floor slabs slid horizontally relative to the supporting beam and helped

to mitigate spalling of the concrete ledge.

3. The timber infill connection utilised between the lateral load resisting frame and

adjoining hollow-core spanning parallel to the frame performed well during the

experiment. This connection detail allowed the flooring and frame elements to deform

freely. At the same time, the integrity of the connection, in terms of the capability of

transferring inertial forces from the floor to the frame, was maintained. The good

performance of the timber infill detail supports its inclusion into New Zealand Concrete

Design Standards. The displacement incompatibility between the floor and frame was

measured during the experiment and correlations between these results and analytical

modelling were good.

4. The lateral load capacity of the super-assemblage was able to be predicted with

satisfactory agreement observed with the experimental results. Although it was seen that

the moment resisting frame dominated the overall super-assemblage strength, the

contribution of the floor slab was noticeable. Compression forces induced in the

longitudinal frame beams due to the bowstring effect were quantified and the progressive

activation of starter bars within the transverse beams was accounted for in calculations of

the negative moment capacities of the frame beams.

Page 5-3
5. As a result of the rigidly reinforced connection between the transverse supporting beams

and floor slabs, torsional hinging at the ends of the transverse beams was observed. The

strong floor-to-beam connection caused the floor to remain horizontal and the supporting

beams to stay essentially upright throughout the cyclic lateral loading of the super-

assemblage in the longitudinal direction, and this placed large torsional demands on the

ends of the transverse beams. It was seen that the flexural hinging of the transverse

beams during transverse loading caused deterioration of the ends of the transverse

beams, which accelerated the effects of torsion observed under longitudinal re-loading.

6. Sub-assemblage two-dimensional proof of concept testing conducted preceding the full-

scale super-assemblage test was seen to provide an accurate representation of the three-

dimensional test. A new testing procedure was developed to include the dual sources of

damage of the relative rotation between the supporting beam and slab, and tension forces

in the floor. Although the limitations of the two-dimensional form of test were

acknowledged, the ability to achieve useful and realistic results for less time and

resource constraints make this format of testing a valuable method of investigation floor-

to-beam connections.

7. This overall experimental investigation has shown that a correct level of reinforcement

between the floor slab and supporting beam can lead to good performance. However,

consideration of previous sub-assemblage research, and the relative performance of each

of the sub-assemblage component specimens tested during this current research indicate

that over-reinforced beam-to-floor connections can cause inferior performance. It is

therefore imperative that a correct level of reinforcement, such as the reinforcement used

for the beam-to-floor slab connection in the current super-assemblage, is used.

Page 5-4
5.3 RECOMMENDATIONS FOR PROFESSIONAL PRACTICE

One of the main objectives of this research was to validate previously unproven and untested

details proposed in amendments to the New Zealand Concrete Design Standard

(NZS 3101, 2004). Recommendations from previous research (Matthews, 2004 and Lindsay,

2004) have also been incorporated to enhance the seismic performance. The following

recommendations have been made for improving professional practice and design standards:

1. Subject to the good performance of the reinforced connection detail used between the

hollow-core floor slabs and supporting beams, it is recommended that this detail is fully

adopted by New Zealand Concrete Design Standards. Figure 5-1 shows a diagram of the

connection detail. Also it can be recommended that the current 1.2% drift design limit

imposed on this class of construction be lifted.

D12 Starter
Bars (Grade 300) 900mm (3*hollowcore depth) D12 Bars @ 300mm each
(300mm spacing) way as topping reinforcing
600mm Lap

300 Series Hollow-core

2*R16 (Grade 300) - 1 per filled core


Low Friction
bearing strip Note: Alternate (every second)
Seat Transverse cores filled and reinforced
Reinforcement 75mm seat (minimum)

Figure 5-1 Recommended hollow-core to supporting beam connection detail.

2. For cases when an infill slab is used to connect an exterior concrete frame and parallel

spanning precast floor, the detail presented in Figure 5-2 should be used. The starter bars

from the perimeter beams should extend a minimum distance of 600mm into the first

floor unit.

Page 5-5
75mm Concrete Topping
D12 Starters extending 600mm
into the first hollow-core unit Timber in-fill
(lapped with diaphragm reinforcing) 750mm

Hollowcore Units
200mm*25mm Timber boards

Perimeter beam

Figure 5-2 Hollow-core connection detail with a parallel load resisting frame

3. Conventional reinforcement should be used in place of customary cold drawn or ductile

mesh in diaphragm topping slabs. Grade 300 D10 or D12 reinforcement (the D

indicates deformed reinforcement) placed at 300mm centres provide sufficient ductility,

decrease the risk of reinforcment fracture and are compatible with conventionally used

Grade 300 D12 starter bar reinforcement.

4. Low friction bearing strips should be used for the seating of hollow-core floor units.

These allow horizontal sliding of the floor units relative to the supporting beam and help

distribute stresses reducing the chance of concrete spalling from the supporting ledge.

Figure 5-3 shows the layout that the seating of precast units should adhere to.

Floor Unit

Supporting Beam
Low friction
bearing strip

Concrete Seat (ledge)

10mm 15mm
50mm

75mm
Figure 5-3 Precast concrete floor seating requirements (close-up view)

Page 5-6
5. Hollow-core units should not be seated on potential plastic hinge zones of supporting

beams. Infill slabs should be used over these regions and additional hooked

reinforcement should be cast in the transverse beams, such as is shown in Figure 5-4, to

force plastic hinging of the supporting beams to occur away from the seated hollow-core

floor units.
Starter bars from the
longitudinal beams Column
(extended into first hollowcore unit)

Hollowcore Units
Timber Infill

Transverse Beam
Hooked Bar
(in addition to the regular reinforcing) Longitudinal Beam

Figure 5-4 Recommended floor supporting beam detailing

5.4 RECOMMENDATIONS FOR FUTURE RESEARCH

Although this research has addressed several issues that have arisen from previous research

findings, there are still many issues that require research attention. The following lists some

areas that need research consideration:

1. Despite the limitations, the effectiveness and ease of conducting proof of concept sub-

assemblage tests relative to full-scale super-assemblage experiments mean that several

other beam-to-floor connections should be investigated. Research should strive towards

testing all floor-to-beam connections that have been used in New Zealand construction,

including other floor types. This research would be able to highlight vulnerabilities

possessed by other forms of construction and help direct further investigations.

Page 5-7
2. The research conducted in this thesis as well as Lindsay (2004) has successfully

provided validated design solutions for future construction. However the shortcomings

of hollow-core flooring observed during the Matthews (2004) testing indicate that many

existing buildings would be damaged in a seismic event. Future research should look to

find suitable economical and practical retrofit measures for existing structures.

3. As the timber infill detail has been proven as a suitable detail to connect lateral frames

with parallel spanning precast floors, future research should seek to develop a flat

soffit option where the flooring is directly adjacent to the perimeter frame. Research has

previously been conducted on such connections, however more specific investigations

need to be undertaken focusing on New Zealand construction conditions. The use of

discrete floor-to-frame connectors that are able to allow vertical displacements but

dissipate energy and transfer inertial loads is one example of a possible connection

solution. A suitable flat soffit detail would offer designers another option should an

infill detail be architecturally undesirable.

4. As the Matthews (2004) and Lindsay (2004) experiments have shown, hollow-core floor

units are susceptible to damage when seated on plastic hinge zones of their supporting

beams. Future research should seek to develop advanced seating materials that allow

both the horizontal sliding of the floor units relative to the supporting beams and

deformations of the supporting beams to be accommodated. Such solutions as rubber

bearings should be investigated in more depth.

Page 5-8
In addition to the experimental research recommended, there is also the need for analytical

work in order to better predict and assess the behaviour of hollow-core structures. Ideally, all

experimental research should be coupled with parallel analytical work so that experimental

work can be better explained and understood as well as to offer a useful tool for researchers

and designers. More specifically, areas of analytical research should develop what has been

touched on in this thesis. Modelling of torsion, the contribution of strength from the floor

slabs in terms of the bowstring effect and displacement incompatibility all need to be

extended.

5.5 REFERENCES

Lindsay R.A, 2004, Experiments on the seismic performance of hollow-core floor systems in

precast concrete building, Masters Thesis, Department of Civil Engineering,

University of Canterbury, Christchurch, New Zealand.

Matthews J.G, 2004, Hollow-core floor slab performance following a severe earthquake, PhD

Thesis, Department of Civil Engineering, University of Canterbury, Christchurch,

New Zealand.

NZS3101, 1995, Concrete Structures Standard, NZS3101, Parts 1 & 2, Standards New

Zealand, Wellington, New Zealand.

NZS3101, 2004, Amendment No. 3 to 1995 Standard (NZS3101), Standards New Zealand,

Wellington, New Zealand.

Page 5-9
Appendix A Sub-Assemblage Experiments: Construction &
Results

A.1 CONSTRUCTION DETAILS AND PHOTOGRAPHS

Full size
test SFD based on
12m span
35kN prototype

SFD based on
6m span

SFD due to additional


mass (Pt load = 35KN)

Combined SFD for


SW + Pt load

Matching the Comparison between the SFD


shears at the for both the 12m unit and the
beam connection 6m unit with an additional mass

Reaction provided by vertical


actuator

(a) Schematic view of the link with the super-assemblage

Reaction Frame

Reaction Frame

35kN Concrete Vertical


Weight (drift controlling)
Actuator

Support beam and


restraints

Hollow-core unit
300 series hollow-core unit

6000mm Horizontal
Support beam and
(beam elongation)
restraints
Actuator

South (side) elevation West (end) elevation

(b) Test loading set-up

Figure A-1 Test set-up and loading configuration.

Page A-1
(a) Supporting beam details
Floor Unit

D12 Starter
Bars (Grade 300)
(300mm spacing) Supporting Beam
Low friction
bearing strip

Concrete Seat (ledge)

2*R16 (Grade 300) - 1 per filled core 10mm 15mm


Low Friction 50mm
bearing strip Note: Alternate (every second) 75mm
Seat Transverse cores filled and reinforced
Reinforcement 75mm seat (minimum)

(b) Specimen 1 beam-to-slab connection details


D12 Starter Bars (Grade 500) @ 300mm,
continuous as topping reinforcing

4*D12 (Grade 500) - 2 per filled core


Seat Transverse Low Friction
Reinforcement bearing strip
75mm seat

(c) Specimen 2 beam-to-slab connection details

Figure A.2 Sub-assemblage Specimen design details

Page A-2
(a) Placement of the low friction bearing strip (b) Cutting of the hollow-core cells.
on the concrete seat.

(c) Supporting beam cast in restraints. (d) Formwork prior to concrete casting.

(e) End view of the loading connections. (f) Top view after concrete casting.

Figure A-3 Specimen 1 construction photographs

Page A-3
(a) De-bonding of the reinforcing. (b) View of the supporting beam reinforcement
and anchorage of the starter and hairpin bars.

(c) Top view of the reinforcement in supporting (d) Reinforcement details for the connection
beam and topping slab prior to concrete with the loading actuators at the end of the
casting. hollow-core unit.

(e) Top view of the test specimen before testing.

Figure A-4 Specimen 2 construction photographs

Page A-4
A.2 MATERIAL TESTS

Tensile tests were conducted to determine the yield and ultimate strengths of the provided

concrete. A summary of the tensile test results are shown in Table A-1, a photograph showing

reinforcement test coupons are shown in Figure A-5 and stress strain relationships from the

tensile tests are shown in Figure A-6.

Readymix concrete was used for the casting of the sub-assemblage elements. Standard

30MPa structural concrete mixes were specified with a 120mm slump and 19mm aggregate.

Table A-1 also includes the measured compression strengths ascertained by compression

tests.

Table A-1 Material test properties

Concrete Compression Test Strengths (MPa)


Location 7 Day 28 Day Test Day
Supporting Beams 26.7 37.8 45.3
Specimen 1 Topping Concrete 21.2 30.1 41.2
Specimen 2 Topping Concrete 25.1 32.0 35.2

Reinforcement Tensile Test Strengths


Location Reinforcing fy (MPa) fu (MPa)
Supporting Beams XD25 536 675
Specimen 1: Starters D12 311 450
Specimen 1: Hairpins R16 325 469
Specimen 2: Starters & Hairpins XD12 570 709

Figure A-5 Reinforcement test samples

Page A-5
500 500
450 450
400 400
350 350
Stress (MPa)

Stress (MPa)
300 300
250 250
200 200
150 150
100 100
50 50
0 0
0 0.05 0.1 0.15 0.2 0.25 0.3 0 0.05 0.1 0.15 0.2 0.25 0.3
Strain Strain

(a) Specimen 1: D12 reinforcement. (b) Specimen 1: R16 reinforcement.


800 800

700 700

600 600

Stress (MPa)
Stress (MPa)

500 500

400 400

300 300

200 200

100 100

0 0
0 0.05 0.1 0.15 0 0.05 0.1 0.15 0.2
Strain Strain

(c) Specimen 2: XD12 reinforcement. (d) XD25 Supporting beam reinforcement.

Figure A-6 Reinforcement test properties

A.3 TESTING PHOTOGRAPHIC LOG

A.3.1 Experiment Set-up and Instrumentation

(a) Supporting beam (b) Actuator set-up (c) Actuator set-up


instrumentation
Figure A-7 Instrumentation and loading set-up.

Page A-6
A.3.2 Specimen 1

(a) Continuity crack at the (b) Vertical crack between the (c) Top view of the interface
beam-slab interface at 0.5%. slab and beam at +0.5%. crack at +1%.

(d) Topping cracks at 2%. (e) Widening of the beam-slab interface crack
and exposure of the starter bars.

(f) Interface crack at 3%. (g) Damage at the end of testing.

Figure A-8 Specimen 1 testing photographs

Page A-7
A.3.3 Specimen 2

(a) Continuity crack at the (b) Vertical crack between the (c) Vertical crack between the
beam-slab interface at 1%. slab and beam at -2%. slab and beam at +3%.

(d) Initial signs of concrete (e) Widening of the beam-slab (e) Large opening of the
seat spalling during the 3% interface crack at +4%. beam-slab interface crack at
cycles. -5% drift.

(f) Progression of spalling observed during the (g) Deterioration of the seat and exposure of
4% cycles. the bearing strip.
Figure A-9 Specimen 2 testing photographs

Page A-8
(a) Vertical dropping of the hollow-core floor (b) Hollow-core unit sliding off and down the
slab during the 5% cycles. damaged seat concrete.

(c) Damage to the topping concrete and extent (d) Failure of the test specimen with a loss of
of vertical dropping during the 5% cycles. seating occurring.

Figure A-10 Specimen 2 photographs during the later stages of testing.

Page A-9
A.4 LOADING PROTOCOL

The following table provides a list of the loading input files for the sub-assemblage testing.

For each increment of vertical drift displacement, the calculated elongation is shown, and the

actual actuator target displacements were listed. Figure A-11 shows the interaction between

the vertical and horizontal actuators.

Horizontal
Vertical Ram Ram
Half Cycle Drift Amplitude Drift Amplitude Elongation Displacement Displacement
Count % rads mm mm mm
1 0 0 0.00 0.0 0.0
2 0.5 0.005 2.63 -28.5 -1.9
3 0 0 0.00 0.0 0.0
4 -0.5 -0.005 2.63 28.5 -2.8
5 0 0 0.00 0.0 0.0
6 0.5 0.005 2.63 -28.5 -1.9
7 0 0 0.00 0.0 0.0
8 -0.5 -0.005 2.63 28.5 -2.8
9 0 0 0.00 0.0 0.0
10 1 0.01 5.25 -57.0 -3.2
11 0 0 2.63 0.0 -2.6
12 -1 -0.01 7.88 57.0 -7.5
13 0 0 5.25 0.0 -5.2
14 1 0.01 7.88 -57.0 -5.8
15 0 0 5.25 0.0 -5.2
16 -1 -0.01 7.88 57.0 -7.5
17 0 0 5.25 0.0 -5.2
18 2 0.02 13.13 -114.0 -6.7
19 0 0 10.50 0.0 -10.5
20 -2 -0.02 18.38 114.0 -15.3
21 0 0 0.00 0.0 0.0
22 2 0.02 18.38 -114.0 -11.9
23 0 0 15.75 0.0 -15.7
24 -2 -0.02 18.38 114.0 -15.3
25 0 0 15.75 0.0 -15.7
26 3 0.03 23.63 171.0 -10.3
27 0 0 21.00 0.0 -21.0
28 -3 -0.03 28.88 -171.0 -20.6
29 0 0 26.25 0.0 -26.2
30 3 0.03 28.88 171.0 -15.5
31 0 0 26.25 0.0 -26.2
32 -3 -0.03 28.88 -171.0 -20.6
33 0 0 26.25 0.0 -26.2
34 4 0.04 34.07 228.0 -11.5
35 0 0 31.51 0.0 -31.5
36 -4 -0.04 39.32 -228.0 -23.5
37 0 0 36.76 0.0 -36.8
38 4 0.04 36.76 228.0 -36.8
39 0 0 39.37 0.0 -16.7
40 -4 -0.04 36.76 -228.0 -36.7
41 0 0 39.37 0.0 -23.6

Page A-10
300

200
Vertical Ram Displacement Control (mm)

100

0
-40 -35 -30 -25 -20 -15 -10 -5 0

-100

-200

-300
Horizontal Ram Displacement Control (mm)

Figure A-11 Input loading file interaction between the horizontal and vertical actuators.
(Note that negative displacements denotes retraction of the actuator)

A.5 LOW CYCLE FATIGUE PREDICTION

A low cycle fatigue theory proposed by Dutta and Mander (2001), enables the plastic rotation

life capacity of reinforcement to be calculated. The theory is expressed as follows:

D L p 1 L p
Tp 0.16 1  2 N f
 0.5
(A.1)
D' D 2 L
where Tp = plastic drift; L = lever arm of the cantilever column; D = overall member depth;

D = distance between outer layers of steel, 2Nf = the number of reversals to the appearance of

the first fatigue crack with which Nf is the effective number of constant amplitude cycles for a

variable amplitude displacement history and Lp = equivalent plastic hinge length.

Miners well-known cycle counting method can be utilised to calculate the equivalent

number of equi-amplitude cycles (Nf):

Page A-11
2
T
Nf i T

(A.2)
ref

Where T = drift for the ith cycle of loading and Tref = reference drift for an equivalent

constant amplitude loading history.

The following calculations show how the solution was obtained for this case:

T
2
ncycles (T/Tref)
D= 375 mm 0.5 2 0.03125
D' = 280 mm 1 2 0.125
Lp = 150 mm 2 2 0.5
L= 6000 mm 3 2 1.125
Tref = 4 % 4 2 2
Ty = 0.7 % 6 3.7813

375 150 1 150


2 u 3.78
 0.5
T Ty T p 0.007  0.16 1  0.037 rads
280 375 2 6000

This compares well with the experimental observations, where the first fracture occurred at

3.83%.

A.6 REFERENCES

Dutta A and Mander J.B, 2001, Energy Based Methodology for Ductile Design of Concrete

Columns, Journal of Structural Engineering, Vol. 127(12), December, pp 1374-1381.

Page A-12
Appendix B Super-Assemblage Construction Photographs and
Design Details

B.1 BUILDING CONSTRUCTION DRAWINGS

The following pages display the construction drawings of the super-assemblage frame

structure:

B-2 South-West Beam-Column Joint and Beams Precast Unit Plan

B-3 South-West Beam-Column Joint and Beams Precast Unit Elevation

B-4 South-East Beam-Column Joint and Beams Precast Unit Plan

B-5 South-East Beam-Column Joint and Beams Precast Unit Elevation

B-6 North-West Beam-Column Joint and Beams Precast Unit Plan and Elevation

B-7 North-East Beam-Column Joint and Beams Precast Unit Plan and Elevation

B-8 South-Centre Beam-Column Joint and Beams Precast Unit Plan and Elevation

B-9 Beam Cross Sections

B-10 South-East and South-West Column Top Segments

B-11 South-East and South-West Column Bottom Segments

B-12 North-East and North-West Column Top Segments

B-13 North-East and North-West Column Bottom Segments

B-14 South-Centre Column Top Segment

B-15 South-Centre Column Bottom Segment

B-16 Longitudinal Beams Lap Splice Details

B-17 Transverse Beams Lap Splice Details

B-18 Lap Splice Beam Cross Sections

B-19 Back Tie Frame Details

Page B-1
Page B-2
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury Precast Unit Details Date: 27/11/03
Christchurch, New Zealand SW Unit Plan Revision: A PU1
Scale: 1:25
Page B-3
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury Precast Unit Details Date: 27/11/03
Christchurch, New Zealand Revision: A PU2
SW Unit Elevations Scale: 1:25
Page B-4
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury Precast Unit Details Date: 27/11/03
Christchurch, New Zealand Revision: A PU3
SE Unit Plan Scale: 1:25
Page B-5
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury Precast Unit Details Date: 27/11/03
Christchurch, New Zealand Revision: A PU4
SE Unit Elevations Scale: 1:25
Page B-6
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury Precast Unit Details Date: 27/11/03
Christchurch, New Zealand Revision: A PU5
NW Unit Plan & Elevation Scale: 1:25
Page B-7
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury Precast Unit Details Date: 27/11/03
Christchurch, New Zealand Revision: A PU6
NE Unit Plan & Elevation Scale: 1:25
Page B-8
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury Precast Unit Details Date: 27/11/03
Christchurch, New Zealand Revision: A PU7
SC Unit Plan & Elevation Scale: 1:25
Page B-9
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury Precast Unit Details Date: 27/11/03
Christchurch, New Zealand Revision: A PU8
Beam Cross Section Details Scale: 1:20
Page B-10
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury
Precast Unit Details Date: 27/11/03
Christchurch, New Zealand SE & SW Column Tops Revision: A PU9
Scale: 1:25
Page B-11
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury Precast Unit Details Date: 27/11/03
Christchurch, New Zealand Revision: A PU10
SE & SW Column Bottoms Scale: 1:25
Page B-12
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury
Precast Unit Details Date: 27/11/03
Christchurch, New Zealand NE & NW Column Tops Revision: A PU11
Scale: 1:25
Page B-13
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury Precast Unit Details Date: 27/11/03
Christchurch, New Zealand Revision: A PU12
NE & NW Column Bottoms Scale: 1:25
Page B-14
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury
Precast Unit Details Date: 27/11/03
Christchurch, New Zealand SC Column Top Revision: A PU13
Scale: 1:25
Page B-15
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury Precast Unit Details Date: 27/11/03
Christchurch, New Zealand Revision: A PU14
SC Column Bottom Scale: 1:25
Page B-16
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury
In-Situ Beam Lap Splice Details Date: 27/11/03
Christchurch, New Zealand Longitudional Beams (Beams 1 & 3) Revision: A BS1
Scale: 1:25
Page B-17
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury
In-Situ Beam Lap Splice Details Date: 27/11/03
Christchurch, New Zealand Transverse Beams (Beams 2 & 4) Revision: A BS2
Scale: 1:25
Page B-18
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury Beam Lap Splice Date: 27/11/03
Christchurch, New Zealand Revision: A BS3
Cross Section Details Scale: 1:20
Page B-19
Department of Civil Engineering Drawn: CJM Drawing No.
University of Canterbury Back Frame Date: 01/03/04
Christchurch, New Zealand Revision: A BF1
Tie Beam Details Scale: 1:50
B.2 BUILDING CONSTRUCTION PHOTOGRAPHIC LOG

(a) North-west column base (b) Column bases prior to (c) Mid-span splice of east
reinforcement cage casting transverse beam

(c) South east beam column joint (d) Presence of ducting to allow connections of
reinforcement the load frame brackets

(e) Pouring of the column bases (f) Extrusion of the hollow-core units

Figure B-1 Construction of the precast components.

Page B-20
(a) Lowering of the south-east precast unit (b) Use of a mobile crane for erecting the
precast frame

(c) Moving the north-west precast unit into (d) The frame in place
place

(e) Arrival of the hollow-core floor units (f) Floor units in place

Figure B-2 Erecting of the precast units and placement of the hollow-core floor units.

Page B-21
(a) Grouting of the column reinforcement (b) View of the starter bar reinforcement and
ducting and dry packing the precast interface backing board coverage of the un-filled cores

(c) Placement of the low (d) The supporting beam (e) View of the central column
friction bearing strip on the reinforcement darg bar reinforcement
concrete seat

(e) Topping reinforcement over the timber (f) Diaphragm topping concrete
infill
Figure B-3 Reinforcement detailing of the supporting beams and topping slab.

Page B-22
(a) Concrete pumping (b) Pouring of the concrete topping

(c) Vibration of the poured concrete (d) Manual floating and levelling of the
topping concrete

(e) Kelly float finishing of the topping concrete (f) Curing of the topping concrete

Figure B-4 Pouring of the topping slab.

Page B-23
(a) Laboratory before super-assemblage (b) Laboratory after super-assemblage
construction construction

(c) Completed and painted super-assemblage (d) Back reaction frame and tie beam set-up
structure

Figure B-5 Before and after super-assemblage photographs

Page B-24
Appendix C Super-Assemblage Testing and Photographic Log

C.1 MATERIAL TESTING

C.1.1 Reinforcement Testing

New Zealand manufactured reinforcement was used for this experimental project. Table C-1

presents the reinforcing properties obtained from the tensile testing carried out using the

Avery testing machine in the Department of Civil Engineering Laboratory. Figure C-1 shows

the stress strain relationships obtained from the tensile tests.

Table C-1 Reinforcement tensile test properties.

Tensile Test Strength and Strain Characteristics


Location Reinforcing fy (MPa) fsu (MPa) Es (GPa) Esh (GPa) Hsh Hsu
Column & Beam Longitudional Steel YD25 536 675 214 3.27 0.0015 0.0128

Column & Beam Transverse Steel YR12 530 714 190 - 0.0005 0.0152
Support Connection R16 336 469 186 4.48 0.0031 0.0200
Starter Bars and Diaphragm D12 307 447 181 3.65 0.0029 0.0218
Drag Bars YD20 580 724 246 3.82 0.0012 0.0156

Page C-1
500 500
450 450
400 400

350 350

Stress (MPa)
Stress (MPa)

300 300

250 250

200 200

150 150

100 100

50 50

0 0
0 0.05 0.1 0.15 0.2 0.25 0.3 0 0.05 0.1 0.15 0.2 0.25 0.3
Strain Strain

(a) D12 starter and topping reinforcement (b) R16 hairpin reinforcement
800
800
700
700
600
600
500
Stress (MPa)

500
Stress (MPa)

400 400
300 300
200 200

100 100

0 0
0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.2
Strain Strain

(c) XD20 drag bar reinforcement (d) XR12 transverse reinforcement

800

700

600

500
Stress (MPa)

400

300

200

100

0
0 0.05 0.1 0.15 0.2
Strain

(e) XD25 beam and column reinforcement


Figure C-1 Reinforcement stress strain tensile test results

Page C-2
C.1.2 Concrete Testing

Standard ready mix concrete was supplied for this project. The mixes were specified with a

30MPa 28-day compressive strength, 19mm maximum aggregate size and 100mm slump. For

the topping slab pour, a pump mix was used with a specified 28-day compression strength of

30MPa and 13mm maximum aggregate size. Compression tests were undertaken at 7-day, 28-

day and the initial day of testing. Table C-2 shows a summary of the compression test data

and Figure C-2 shows a photograph of preparation of the concrete cylinders.

Table C-2 Concrete compression strengths.

Compression Test Strengths (MPa)


Location 7 Day 28 Day Test Day
Precast Frame Units 1 26.7 37.8 45.3
Precast Frame Units 2 21.2 30.1 41.2
Precast Frame Units 3 25.1 32.0 35.2
Splice Half Beams 31.8 - 46.0
Topping Slab & Half Beams 31.0 - 40.6
Grout: Column Reinforcing - - > 60

Figure C-2 Preparation of concrete test cylinders.

Page C-3
C.1.3 Hollow-core Properties

Hollow-core concrete units are cast on long casting bed in a single pour and cut to length

following curing and hardening of the concrete. Prestressing strands are tensioned prior to an

extrusion machine travelling along the casting bed extruding the hollow-core units using

zero-slump concrete. Several different types of hollow-core units are available which vary in

depth, with 200, 300 and 400mm depths available, and number of prestressing strands. A

cross section of a typical unit is shown in Figure C-3 and common hollow-core properties are

presented in Table C-3.

Figure C-3 Typical 300 series hollow-core unit cross-section.

Table C-3 Section properties of a 300 series hollow-core unit.


Hollow-core Section Properties
2 4
Section Area (m ) yb (mm) I (m ) Self Weight (kPa) 28-day f'c (MPa)
-3
300 Dycore / Partek 0.1606 153 2.04*10 3.5 42

Page C-4
C.2 INSTRUMENTATION PHOTOGRAPHS

(a) Inclinometers to measure (b) Potentiometers under the (c) Vertical levelling of the
the column drift rotations. timber infill. floor surface.

(d) Inclinometer and potentiometer (e) Potentiometers measuring floor pull-off.


instrumentation of the west transverse beam.

(f) Sonic transducers measuring the lateral (g) Actuator at base of north west column with
(north-south and east-west) displacements of load-cell and rotary potentiometer.
the columns.
Figure C-4 Instrumentation photographs

Page C-5
(a) Instrumentation of the south-west (b) Instrumentation of the south-centre beam
longitudinal beam and beam column joint. column joint.

(c) View of the computers used for controlling (d) Data logging and controller boxes.
the experiment and data acquisition.

(e) Load frame set-up for longitudinal loading. (f) Load frame set-up for transverse loading.
Figure C-5 Instrumentation and load-frame photographs

Page C-6
C.3 TESTING PHOTOGRAPHIC LOG

C.3.1 Phase 1: Longitudinal loading

(a) Cracking of the infill slab at +0.5% (b) Cracking of the transverse beam (north
east end) at +0.5%

(c) Diagonal cracking within the south centre (d) Continuity crack looking towards the north
beam column joint at 0.5% west column at 0.5%

(e) Cracking over the infill (f) Continuity crack at the (g) Diagonal torsion cracks,
slab at 1% west end at 1% north west corner at 1%
Figure C-6 Phase 1: Damage during the early stages

Page C-7
(a) Cracks extending over the timber infill slab (b) Hollow-core cracking at the south east
(south west bay) at +2% corner at +2%

(c) Web crack forming at south east corner (d) First instance of spalling noticed on east
unit at +2% transverse beam at -2%

(e) Continuity crack along the (f) Crack opening at the south (g) Longitudinal topping
beam-slab interface at the east column at +2% crack between 1st and 2nd
eastern end at +2% hollow-core unit
Figure C-7 Phase 1: Damage during the later stages

Page C-8
C.3.2 Phase 2: Transverse loading

(a) Spalling and exposure of the bearing strip (b) 13mm crack opening at the north east
of the east transverse beam at +1% column at +2%

(c) Large crack at the beam-column interface (d) Top view of the damage at the north east
of the south east corner at +2% corner of the structure at 3%

(e) Damage to the south west (f) Damage close to the (g) South east column
transverse hinge at +3% column face at the north east inclination at 3%
transverse beam end (+3%)
Figure C-8 Phase 2 Transverse loading damage

Page C-9
C.3.3 Phase 3: Longitudinal re-loading

(a) Distribution of topping cracking over the (b) Large crack opening at the south east
south east bay at 2% column at +3%

(c) View of the damage to the south east south (d) Exposure of the reinforcement and vertical
frame hinge at +3% dropping of the west transverse beam (north
end) at +3%

(e) Hollow-core soffit cracking (f) Damage due to torsional (g) Torsional hinging looking
at the east end at +3% hinging at +3% (south east) towards the north east column
Figure C-9 Phase 3: Damage during the early stages

Page C-10
(a) Crushing at the top west (b) Torsional behaviour of the (c) Interface cracking of the
side of the south centre east transverse beam at 3% south east hinge of the south
column at 3% beam at +4%

(d) West transverse beam (e) Damage of the north east (f) Damage of the north west
remaining upright at +4% hinge (east beam) at +4% hinge (west beam) at +4%

(g) Buckling of the bottom reinforcement of the (h) Top view of the damage at the south west
south west beam west hinge at 4% corner of the floor at 4%
Figure C-10 Phase 3: Damage during the middle stages

Page C-11
(a) Damage to the north east (b) Hollow-core soffit (c) Different rotations of the
hinge and evidence of the cracking, 2nd unit, west end at east beam and columns due to
beam dropping at 4% 4% torsion at +5%

(d) The plastic hinges (from west to east and left to right) of the south frame at +5% drift

(e) Evidence of hollow-core pull-off (un- (f) Increasing buckling of the reinforcement of
painted soffit) at +5% the south west beam hinge at 5%

Figure C-11 Phase 3: Damage during the later stages

Page C-12
C.3.4 End of Testing

(a) Fracture of the top, outside bar of the south (b) Fracture of the inside and outside top bars
east hinge of the south beams of the south east hinge of the south beams

(c) Cracking over the timber infill of the south east bay at the conclusion of testing

(d) Cracking over the timber infill of the south west bay at the conclusion of testing

Figure C-12 Notable damage following the conclusion of testing.

Page C-13
C.4 CONCRETE TOPPING DELAMINATION

The occurrence of the concrete topping slab (75mm thick) delaminating from the hollow-core

floor units was recorded throughout testing. A hammer was used to find hollow sounding

areas indicative of delamination. As Figure C-12 shows, delamination was initially noticed at

the 2% peaks during Phase 1, and area of delamination gradually increased throughout the

remainder of testing. Delamination extended along the length of the building at the infill-to-

Unit 1 interface, along the beam-to-floor slab interfaces at the ends, and in the southern

corners at the ends of Unit 1.

Unit 4

Unit 3

Unit 2

Unit 1

Timber Infill

Key (note that the delamination is cumulative)

After +2%, Phase 1


After -2%, Phase 1
After Phase 2
After Phase 3

Figure C-13 Topping slab delamination pattern

Page C-14
Appendix D Analytical Results

D.1 YIELD DRIFT CALCULATION

It is difficult to accurately determine the experimental yield drift, as there is no clear change

in stiffness of the super-assemblage, and it was seen that yielding occurred between around

0.5% drift and 0.7% drift. A yield value of 0.5% drift was used for the formulation of beam

elongation results. The following calculations show the contributions of the beam, column

and beam-column joint towards the calculated yield drift rotation. The calculations follow the

same steps as used by Matthews (2004) and Lindsay (2004).

D.1.1 Beam Contribution

The following relationship from Priestley and Kowalsky (2000) was utilised to calculate the

contribution from the beam. The yield moment was derived from simple mechanics and actual

material test results;

Hy
I yb 1.70 6.06 u 10 6 (D-1)
hb

where: Iyb = yield curvature of the beam, Hy = yield strain of the beam longitudinal

reinforcement and hb = height of the beam. Knowing the nominal moment capacity of the

beam and the previously calculated yield curvature of the beam, the effective stiffness can be

determined;

M yb 498 u 10 6
EI eff 8.22 u 1013 ( Nmm 2 ) (D-2)
I yb 6.06 u 10 6

where: (,eff = effective flexural stiffness and 0yb = nominal yield moment capacity of the

beam section. Using moment area theorem, the deflection of the beam can be found from the

following relationship:

Page D-1
M yb L2b 498 u 10 6 u 5350 2
Gb 14.45mm (D-3)
12 EI eff 12 u 8.22 u 1013

where: Gb = displacement of the beam at its point of inflection and Lb = distance between the

beam hinges. Finally, using the height of the column (Lc) this displacement can be converted

to an overall column displacement (Gbc);

L2c
G cb Gb 16.58mm (D-4)
0.5 u Lb

D.1.2 Beam-Column Joint Contribution

To calculate the contribution to the yield drift from the beam column joint, an expression

derived by Matthews (2004) was used;

As f y (d  d ' ) L
Jj 38.5 1  0.00136rads (D-5)
BH Ec H Lb

where: Jj = joint distortion, % = width of the column, H = height of the column, (d-d) =

internal lever arm of the beam and L = centreline distance between beam inflection points.

Using this calculated rotation, the contribution of the joint can be found in terms of the

column displacement (Gjc) to be 4.76mm.

D.1.3 Column Contribution

The contribution from the column can be calculated following a similar method used for the

beam. The yield curvature for the column can be determined from the following (Priestly and

Kowalsky, 2000):

Hy
I yc 2.12 7.56 u 10 6 (D-6)
hc

Page D-2
where: Iyc = yield curvature of the column, Hy = yield strain of the column longitudinal

reinforcement and hc = width of the column. The effective stiffness can be found using the

nominal moment capacity of the column, 0yc, and is shown by the following equation:

M yc 892 u 10 6
EI eff 1.18 u 1014 ( Nmm 2 ) (D-7)
I yc 7.56 u 10 6

From this, the column contribution can be found from:

M c l c2 223 u 10 6 u (1750  375) 2


Gc 1.2mm (D-8)
3EI eff 3 u 1.18 u 1014

where: 0c = the column moment at the beam face at yielding of the beam (223kNm) and

lc = distance between the beam hinges.

Thus, with all three constituent contributing sources calculated the overall yield drift

can be found from the following:

G G cb  G cj  G c 16.58  4.76  1.2 22.54mm (D-9)

This displacement corresponds to an inter-storey drift of 0.64%.

This calculated value shows good resemblance with the experimental yielding point. It is

likely that the experimental value is higher due to a small amount of cracking of the specimen

prior to testing as well as the cracking that occurred between the precast construction element

interfaces which occurred early in the test. The cracking would influence the effective

stiffness of the section, and the larger overall displacements of the super-assemblage due to

these cracks would result in the yield drift being higher than predicted.

Page D-3
D.2 SUPER-ASSEMBLAGE STIFFNESS

The initial elastic stiffness of the super-assemblage was calculated using the initial results for

each of the three loading phases, and is shown in Figure D-1. It can be seen that the initial

stiffness of structure during Phase 1 loading was 62MN/m (Figure D-1(a)). Figure D-2(b)

shows that the initial stiffness for Phase 2 loading was 34.7MN/m, which is a 44% reduction

of the stiffness exhibited at the start of testing. The cracking of the transverse beams, a large

degree due to torsion reducing the stiffness during Phase 1 loading, can explain this. Cracking

of the floor diaphragm over the timber infill also would reduce the stiffness for the transverse

loading. The stiffness from the early cycles during Phase 3 was 17.5MN/m (Figure D-1(c)).

This is only 28% of the Phase 1 stiffness and is due to the damage accrued from Phase 1 and

Phase 2 loading. The damage of the ends of the transverse beams consequently led to

torsional hinging of these beams, and due to this, the stiffness was significantly reduced. The

widening of the cracks around the north and south corner columns, and continuity crack

between the floor and beams along the east and west bays would reduce the overall super-

assemblage stiffness and decrease the stiffening influence of the floor.

1500 800 2000

600 1500
1000
400 1000
500
200 500
Inter-Storey Inter-Storey Inter-Storey
Drift (mm) Drift (mm) Drift (mm)
0 0 0
-20 -10 0 10 20-30 -20 -10 0 10 20 30-100 -50 0 50 100
-200 -500
-500
Base Shear
Force (kN)

Base Shear
Base Shear

Force (kN)
Force (kN)

-400 -1000
-1000
-600 -1500

-1500 -800 -2000

(a) Phase 1 (b) Phase 2 (c) Phase 3

Figure D-1 Stiffness of the super-assemblage.

Page D-4
D.3 TORSION TEST

A torsion test was undertaken during Phase 2 transverse loading to attempt to determine the

torsional stiffness of the super-assemblage. The test involved displacing the east and west

bays in opposing directions as shown in Figure D-2. For example, positive drifts denote a

positive translation of the east bay (tops of the columns displace north) and negative

translation of the west bay (tops of the columns displace south) and vice versa for the negative

cycle. One reversing cycle to 0.5% was undertaken, and Figure D-3 shows the

corresponding base shear. This was completed after the second cycle to 2%, and a following

third cycle to 2% was undertaken to see if there was any degradation of stiffness due to the

torsion test.

North

Direction Direction
of movement West East of movement

South

Figure D-2 Illustration of the torsion test.

West Bay 300 East Bay


Base Shear

S
Force (kN)

N S N

200

100

Inter-storey Drift (%)


0
-0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 0.6
-100

East Bay -200


West Bay
-300

Figure D-3 Base shear results from the torsion test.

Page D-5
As Figure D-3 shows, the structure had a higher stiffness in positive drift direction, when the

respective tops of the columns translated northwards, than the negative direction. For

example, for the east bay, the stiffness values obtained from the experimental data for the

positive and negative drift cycles were 15.26MN/m and 10.56MN/m respectively, and similar

behaviour was observed for the west bay. The negative drift cycles correspond to negative

moments of the south east and west hinges, and due to the already highly cracked infill slab

section which under negative moments would be in tension, would require less force to reach

the target displacements. This can explain the lower stiffnesses observed during the negative

drift cycles. During the torsion tests, there was also little noticeable damage to the southern

longitudinal beams and after the resumption of the normal testing regime, there was no

noticeable degradation due to the torsion tests.

D.4 GENERAL SUPER-ASSEMBLAGE PERFORMANCE

D.4.1 Central Column Joint Reinforcing Bar Slip

The slipping of the longitudinal top and bottom beam reinforcement passing through the south

central beam column joint was monitored throughout the test by two potentiometers attached

to the reinforcement. Figure D-4 shows the bar slip experimental results for both Phase 1 and

3 longitudinal loading. Figure D-4(a) shows that there was up to 1.9mm of bar slippage

during Phase 1 and during Phase 3 more slipping occurred with a maximum bar slip of

3.82mm being recorded (Figure D-4(b)). The degradation of bond that would allow this bar

slippage would denote a reduction of the stiffness of the super-assemblage during low drift

displacements, which was observed to a small degree. The New Zealand Concrete Design

Standard (NZS 3101:1995) provide design equations to limit the maximum reinforcing size

Page D-6
passing through a beam column joint, which is based on ensuring that sufficient bond is

maintained. These equations are as follows:

db f c'
d 3.3D f (D.10)
hc Do f y

where db = the diameter of bar passing through the joint, hc = height of the column section,

Df = factor to allow for one or two way framing, fc = compression strength of the concrete,

Do = overstrength factor for the beam reinforcement, and fy = yield strength of the

reinforcement. Application of this formula gives a maximum bar diameter of 23.5mm, lower

than the actual XD25 reinforcement used. Although this seems to be an oversight of the

design process, it is a smaller bar diameter than commonly used in large concrete frames and

the New Zealand Concrete Code offers another form of the equation:

db D tD p f c'
d 6 D f (D.11)
hc Ds Do f y

where Dt = factor for the amount of fresh concrete cast below the reinforcement, Dp = factor to

allow for the beneficial effects of axial compression forces, and Ds = factor for severe bond

conditions. In this case, these factors are all 1, and the maximum allowable reinforcement

diameter becomes 42mm, and thus the 25mm bars used are suitable, from a design point of

view, despite the amount of slippage observed.


4 4

3 3

2 2

1 1
Slip (mm)

Scan N umb er
Slip (mm)

Scan N umb er
0 0
0 1000 2000 3000 4000 5000 0 1000 2000 3000 4000 5000 6000
-1 -1

-2 -2
Top
Top
-3 Bottom -3
Bottom
-4 -4
(a) Longitudinal bar slip during Phase 1 (b) Longitudinal bar slip during Phase 3
Figure D-4 Longitudinal beam bar slippage through the central beam column joint

Page D-7
D.4.2 Hollow-core Floor Unit Pull-off

Sprung loaded potentiometers were installed underneath each of the hollow-core units to

measure the horizontal sliding displacements of the floor units relative to the supporting

beam. Figure D-5 shows the relationship between the horizontal pull-off movements and

inter-storey drift. Not surprisingly, and analogous to the beam elongation theoretical

predictions, the amount of pull-off increases with increasing inter-storey drift amplitudes and

is higher for the second cycles. For Phase 1, both the east and west bays showed similar

behaviour, and maximum pull-offs of 12mm and 11.4mm were seen at the second cycle to

2% and +2% for the east and west bays respectively.


14 Displacement

12
(mm)

10

2 East
Drift (%) West
0
-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5

(a) Horizontal hollow-core unit pull-off for Phase 1 loading.


14

12

10

6
Displacement

4
(mm)

East
2
West
Drift (%)
0
-4 -3 -2 -1 0 1 2 3 4
(b) Horizontal hollow-core unit pull-off for Phase 3 loading.

Figure D-5 Horizontal hollow-core unit pull-off

Page D-8
16 14
Unit 1East Unit 1West
14 12
Unit 2 East Unit 2 West
Unit 3 East Unit 3 West
12
Displacement (mm)

Displacement (mm)
10 Unit 4 West
10
8
8
\
6
6
4
4

2 2
Scan N umb er Scan N umb er
0 0
0 1000 2000 3000 4000 0 1000 2000 3000 4000

(a) East beam pull-off: Phase 1 (b) West beam pull-off: Phase 1
12 12

10 10
Displacement (mm)
Displacement (mm)

8 8

6 6

4 4
Unit 1 East Unit 1 West
2 Unit 2 East 2 Unit 2 West
Unit 3 East Unit 3 West
Scan N umb er Unit 4 West Scan N umb er
0 0
0 1000 2000 3000 4000 5000 0 1000 2000 3000 4000 5000

(c) East beam pull-off: Phase 2 (d) West beam pull-off: Phase 2

18 18

16 16
14 14
Displacement (mm)
Displacement (mm)

12 12
10 10
8 8

6 6
Unit 1 West
4 Unit 1 East 4 Unit 2 West
Unit 2 East Unit 3 West
2 Unit 3 East 2 Unit 4 West
Scan N umb er Scan N umb er
0 0
0 2000 4000 6000 8000 0 2000 4000 6000 8000

(d) East beam pull-off: Phase 3 (e) West beam pull-off: Phase 3

Figure D-6 Horizontal hollow-core pull-off displacements.

Figure D-6 shows the horizontal pull-off displacement measured underneath all of the hollow-

core units, related to time, for each of the three loading phases. The fluctuations show the

effect of the cyclic loading, and for Phase 1, the large increases of the pull-off displacement

Page D-9
signify new inter-storey drifts being experienced. The pull-off was seen to be minimal for

Phase 2, and the slight increase of the pull-off, of less than 1mm, can be explained by

torsional behaviour of the transverse beams. The Phase 3 pull-off relationships indicate a

decreased rate of pull-off, as compared with Phase 1 where the Phase 3 contributed a further

4mm to the already 12mm from Phase 1, and the overall maximum pull-off experienced was

approximately 16mm for both bays.

D.4.3 Beam Column Joints and Plastic Hinge Rotation Contributions

Figure D-7 shows a schematic view of the instrumentation used for the beam plastic hinge

and beam-column joint regions. It was hoped that the set-up used would be able to measure

the relative contributions towards the overall rotation, which would equate to the column

inclination. Potentiometers were arranged in such a fashion that joint shear, hinge shear, hinge

flexure, shearing at the beam-column interface and fixed end rotation could be monitored.

Figure D-8 shows the relative components contributing towards the overall lateral

displacement of the structure for Phase 1 and Phase 2 loading.

175mm 700mm
Potentiometers (top and bottom)
Beam-Column Joint Potentiometer measuring the fixed-end rotation
instrumentation - measuring and curvature of the beam plastic
100mm
joint shear deformation 75mm
500mm hinges
100mm

500mm

600mm
600mm

500mm

500mm

600mm Beam Plastic Hinge Potentiometer


300mm 300mm instrumentation - measuring
hinge shear and flexural deformation
195mm

Figure D-7 Close-up representation of the beam-column joint and hinge instrumentation

Page D-10
0.025

Rotation Contribution
0.02
Closure Error

(rads)
0.015

0.01

0.005
Inter-storey Drift (rads)
0
-0.025 -0.02 -0.015 -0.01 -0.005 0 0.005 0.01 0.015 0.02 0.025
-0.005
Fixed End Rotation
-0.01 Hinge Flexure
Hinge Shear
-0.015
Joint Shear
Closure Error -0.02
Column Inclination
Total
-0.025
(a) Contributions towards lateral displacement during longitudinal Phase 1 loading

0.025
Rotation Contribution

0.02 Closure Error


(rads)

0.015

0.01

0.005
Inter-storey Drift (rads)
0
-0.025 -0.02 -0.015 -0.01 -0.005 0 0.005 0.01 0.015 0.02 0.025
-0.005
Fixed End Rotation
-0.01 Hinge Flexure
Hinge Shear
-0.015
Joint Shear
-0.02 Column Inclination
Closure Error Total
-0.025

(b) Contributions towards lateral displacement during transverse Phase 2 loading

Figure D-8 Components that contribute to the overall lateral displacement of the super-
assemblage

By comparing the total rotations with the column inclination in the relationships shown in

Figure D-8, it can be seen that the instrumentation was able to measure the overall lateral

displacements to a satisfactory level. The closure error indicates the discrepancy, and for the

transverse loading as seen in Figure D-8(b), the instrumentation slightly over-predicts the

amount of lateral displacement of the structure. In terms of the individual components, the

Page D-11
fixed end rotation, measured between the column and beam at the top and bottom appeared to

dominate the rotation contributions. For Phase 2, almost all the lateral displacement occurs

from fixed end rotation with a smaller amount due to plastic hinge flexure. This matches well

with the experimental observations whereby damage to the plastic hinges was predominantly

a large crack at the beam-column interface. As it was expected, there were minimal

contributions from the beam hinge and beam column joint shearing displacements.

D.4.4 Reinforcement Strain Information

At several locations within the super-assemblage topping and floor-to-beam connections, the

reinforcement was strain gauged to be able to monitor the demands imposed on the

reinforcement. One main reason was to monitor the strain development along the length of the

starter bars at the end connections to give an idea as to the possible required length of the

starter bars. This meant that strain gauges were attached at several points up to 2.1m from the

supporting beams. Figure D-9 shows the locations of the strain gauges present within the

structure. Figures D-10 and D-11 provide the strain results for the end connection reinforcing

(starter bars, lapped diaphragm steel and the hairpin bars) for the east and west ends during

Phase 1. Figure D-12 shows the results, for Phase 1 and 2, of the southern beam starter bars

that were cast in the infill topping slab and were embedded half way into the southernmost

hollow-core unit. Two starter bars were strain gauged, one 1m either side of the central

column, approximately where the maximum displacement incompatibility between the floor

and beams was expected. Also, the drag bars that were cast into the south central column

and lapped with reinforcement running across the width of the building, were strain gauged at

various points as shown in Figure D-9. Figure D-13 shows the results obtained from the drag

bars for Phase 1 and 2 loading.

Page D-12
North

Unit 4 West East Unit 4


Tie Tie
Bar Bar
Unit 3 Unit 3
West East
Unit 2 Unit 2

Unit 1 Unit 1
West Infill East Infill
Starter Starter

South

(a) Plan view of the location of the strain gauges


Starter bars and diaphragm reinforcement

300 600 600 600

1 3
2 4 5

1 2

300
Hairpin reinforcement

(b) End connection reinforcement strain gauge locations

300mm 375mm
375mm

4 3 2 1

(c) Strain gauging of the starter bars passing over the timber infill

Figure D-9 Strain gauge instrumentation layout

Page D-13
12 45

Interface 40 Interface
10 300mm 300mm
900mm 35 900mm
1500mm 1500mm
8 2100mm 30 2100mm
Yield Strain Yield Strain
strain (%)

strain (%)
25
6
20

4
15

2 10

Scan Num ber 5 Scan N umb er


0
0 1000 2000 3000 4000 5000 0
0 1000 2000 3000 4000 5000
-2 -5

(a) Starter bar strain data for Unit 4 (b) Starter bar strain data for Unit 3
12 18
Interface 300mm 16 Interface
10 900mm 1500mm 300mm
2100mm Yield Strain 14 900mm
1500mm
8 12 2100mm
Yield Strain
strain (%)

6 strain (%)
10

8
4
6
2 4

Scan Num ber 2


0 Scan Num ber
0 1000 2000 3000 4000 5000
0
0 1000 2000 3000 4000 5000
-2
-2

(c) Starter bar strain data for Unit 2 (d) Starter bar strain data for Unit 1
45 45

40 40

35 35

30 30
Unit 2 Interface
Unit 4 Interface Unit 2 300mm
strain (%)
strain (%)

25 Unit 4 300mm 25 Unit 1 Interface


Unit 3 Interface Unit 1 300mm
20 Unit 3 300mm 20 Yield
Yield
15 15

10 10

5 5
Scan Num ber Scan Num ber
0 0
0 1000 2000 3000 4000 5000 0 1000 2000 3000 4000 5000

-5 -5

(e) Hairpin bar strain data, Units 3 & 4 (f) Hairpin bar strain data, Units 1 & 2

Figure D-10 Strain data for the west slab-beam connection reinforcement
during Phase 1

Page D-14
45 45

40 Interface 40 Interface
300mm 300mm
35 900mm 35 900mm
1500mm 1500mm
30 2100mm 30 2100mm
Yield Strain Yield Strain

strain (%)
strain (%)

25 25

20 20

15 15

10
10

5 Scan N umb er
5 Scan N umb er

0
0
0 1000 2000 3000 4000 5000
0 1000 2000 3000 4000 5000
-5
-5

(a) Starter bar strain data for Unit 4 (b) Starter bar strain data for Unit 3
2.5 45

40 Interface
300mm
2 35 900mm
Interface
300mm 1500mm
30 2100mm
900mm Yield Strain
1.5 1500mm
strain (%)
strain (%)

2100mm 25
Yield Strain
1 20

15

0.5 10

5 Scan N umb er
0 Scan Num ber
0 1000 2000 3000 4000 5000 0
0 1000 2000 3000 4000 5000
-5
-0.5

(c) Starter bar strain data for Unit 2 (d) Starter bar strain data for Unit 1
45 45
40 40
35 Unit 4 Interface Unit 2 Interface
Unit 4 300mm 35 Unit 2 300mm
Unit 3 Interface Unit 1 Interface
30 Unit 3 300mm 30 Unit 1 300mm
Yield Yield
strain (%)

strain (%)

25 25
20 20
15 15
10 10
5
Scan Num ber 5
Scan Num ber
0
0 1000 2000 3000 4000 5000 0
0 1000 2000 3000 4000 5000
-5
-5

(e) Hairpin bar strain data, Units 3 & 4 (f) Hairpin bar strain data, Units 1 & 2

Figure D-11 Strain data for the east slab-beam connection reinforcement during Phase 1

Page D-15
30 25
Beam/Infill Beam/Infill
25 Mid-Infill Mid-Infill
20
Infill/Unit 1 Infill/Unit 1
20 Unit 1 Unit 1
Yield Strain 15 Yield Strain
strain (%)

strain (%)
15
10
10
5
5

Scan Num ber 0 Scan Num ber


0 0 1000 2000 3000 4000 5000
0 1000 2000 3000 4000 5000

-5 -5

(a) Strain data for the south west starter bar (b) Strain data for the south east starter bar
(1m west of south central column) during Phase (1m east of south central column) during Phase
1 longitudinal loading 1 longitudinal loading
30 15
Beam/Infill Beam/Infill
20 Mid-Infill 10 Mid-Infill
Infill/Unit 1 Infill/Unit 1
Unit 1 Unit 1
10 Yield Strain 5 Yield Strain
strain
strain

0 0
0 1000 2000 3000 4000 5000 6000 7000 8000 0 1000 2000 3000 4000 5000 6000 7000 8000
Scan Num ber Scan Num ber
-10 -5

-20 -10

-30 -15

(c) Strain data for the south west starter bar (d) Strain data for the south east starter bar
(1m west of south central column) during Phase (1m east of south central column) during Phase
2 transverse loading 2 transverse loading

Figure D-12 Strain data for the starter bars passing over the timber infill

As Figures D-10 and D-11 show, significantly higher strains were experienced at the beam-to-

floor slab interface than in the topping slab, where the strains diminished rapidly. Similar

results were observed for each of the floor units and at both the east and west ends. These

results confirm the both the desired and observed results whereby all of the rotational demand

was centred at the support beam-to-floor interface. The lack of strain penetration would

signify low stresses which was matched by a lack of cracking within the topping away from

the supporting beam. The results for the hairpin bars were less reliable due to malfunctioning

Page D-16
strain gauges, although it can be seen that there was higher strains at the beam-floor interface

than further into the hollow-core units.

Figure D-12(a) and (b) shows that the maximum strains experienced by the south

beam starter bars occurred at the interface between the infill slab and hollow-core floor

Unit 1. This can be explained by the vertical deformations caused by the relative displacement

differences between the floor slab and southern beams and the resultant vertical angle at this

interface. Figure D-13 shows there was a small degree of strain penetration along the drag

bars tying the central column into the floor. Higher strain demand was experienced at the

south centre column. Due to the fact that the southern beam starter bars were also anchored

into the flooring, and that no fracture or tear occurred, the demand on the tie bars was

minimal.

4 6
South Centre Column South Centre Column
3.5 Infill/Unit 1 Infill/Unit 1
5
Unit 1/Unit 2 Unit 1/Unit 2
3 Unit 2/Unit 3 Unit 2/Unit 3
Yield Strain 4 Yield Strain
2.5
strain (%)
strain (%)

3
2

1.5 2

1 1

0.5 Scan Num ber


Scan Num ber 0
0 1000 2000 3000 4000 5000
0
0 1000 2000 3000 4000 5000 -1
-0.5

(a) Strain data for the west drag bar during (b) Strain data for the east drag bar during
Phase 1 longitudinal loading Phase 1 longitudinal loading
6 8
South Centre Column South Centre Column
5 Infill/Unit 1 Infill/Unit 1
Unit 1/Unit 2 6 Unit 1/Unit 2
4 Unit 2/Unit 3 Unit 2/Unit 3
Yield Strain 4 Yield Strain
3
2
strain
strain

1 0
0 1000 2000 3000 4000 5000 6000 7000 8000
0 Scan Num ber
0 1000 2000 3000 4000 5000 6000 7000 8000 -2
-1 Scan Num ber
-4
-2

-3 -6

(c) Strain data for the west drag during Phase 2 (d) Strain data for the east drag during
transverse loading Phase 2 transverse loading
Figure D-13 Strain data for the drag bars tying in the south centre column

Page D-17
D.5 WEST TRANSVERSE BEAM TORSION

D.5.1 Experimental Results

Five inclinometers measuring rotation in the longitudinal direction were spaced along the

west transverse beam at constant intervals as shown in Figure D-14. Figure D-15 shows the

relative torsion rotations between the column and beam for Phase 3, and Figure D-16 provides

the west beam torsion results for Phase 1 and 3.

Figure D-14 Torsion inclinometers attached to the west transverse beam

6
5 Phase 1 Phase 3
Angle of Twist (rad/m)

4
3 Loading Loading
2
1
0
-1
-2
-3
-4
-5
-6
-7
-8

Figure D-15 Torsional twist of the south hinge of the west transverse beam
(radians/metre in plastic hinge zone)

Page D-18
2.5
Jo int
2
1
1.5
Torsion Drift (%)

3
1 4
0.5
S c a n N um be r
0

-0.5 0 500 1000 1500 2000 2500 3000 3500 4000 4500

-1

-1.5

-2

-2.5

(a) West beam torsion during Phase 1


6
Jo int
1
4 2
3
Torsion Drift (%)

4
2
5

S c a n N um be r
0
0 1000 2000 3000 4000 5000 6000 7000 8000

-2

-4

-6

(b) West beam torsion during Phase 3


Figure D-16 Observed torsion of the west beam

D.5.2 Calculation of Torsional Demand

For the present experiment, the torsional demand was caused by the moment from the

connection reinforcement between the supporting beam and floor slabs, which occurs under

lateral loading of the structure. The strength of this connection was found during the sub-

assemblage tests undertaken as part of this research, where calculated moment strengths

matched the experimental values well. The positive and negative moment capacity values

were 38kNm and 40kNm respectively for each hollow-core unit. Accounting for the 1.2m

hollow-core unit widths, the torsional moment per metre length of beam is 32kNm/m and

33kNm/m for the positive and negative rotations, which then corresponds to an applied

torsion at the beam ends of 79kNm approximated for both positive and negative rotations.

Page D-19
For the connection detail tested by Lindsay (2004), the presence of the compressible backing

board meant that there was effectively zero moment resistance. Therefore, the torsional

demands arose due to the eccentric loading of the floor. The weight of hollow-core flooring

was taken to be 3.5kPa, the weight of the topping concrete 24kNm-3 and the eccentricity from

the centre of the beam to the centre of the seat was 238mm. This caused a torsional moment

demand of 20kNm at each end of the transverse beams.

D.5.3 Calculation of Torsional Capacity

For compatibility torsion, as is the case in this experiment, a reduction of the torsional

moment can occur due to the redistribution of internal forces when cracking first occurs. An

expression from the New Zealand Concrete Code, NZS 3101:1995, was used for calculating

the cracking torsional moment of the beam section, and is the same as the formulation given

by Mitchell and Collins (1991). The cracking torsion, or maximum design torsional moment

can hence be calculated by:

f c' Acp2
T *
I (D.12)
3 p cp

in which T* = design torsional moment force at section at the ultimate limit state, or cracking

torsion, I = strength reduction factor (which is omitted in this case for analysis), fc = concrete

compressive strength, Acp = area enclosed by the outside perimeter of the concrete cross-

section, and pcp = outside perimeter of the concrete cross section. Knowing that the overall

concrete cross-section was 400mm wide by 750mm deep, and concrete compressive strengths

obtained from material tests were 35MPa, the torsional cracking moment was 77.2kNm. In

this case, the strength reduction factor was omitted, as this is an analysis of a given section,

not design.

Page D-20
After diagonal cracking has occurred, a truss mechanism of vertical transverse

reinforcing ties and diagonal concrete compression struts forms, and the contribution to the

torsional resistance from the reinforcement can be calculated. Park and Paulay (1975),

provides an expression for the steel contribution to resisting torsion as a function of the

transverse and longitudinal reinforcement, which is shown as Equation D.13. The equivalent

design expression for the nominal torsional strength from NZS 3101:2004 relates to the

amount of transverse reinforcing and stems from information provided by Park and Paulay

(1975).

At f y Al f ly
Ts 2 A0 (D.13)
s P0

in which Ao = area enclosed by the connecting lines between the centres of the longitudinal

reinforcement, At = area of a transverse reinforcing stirrup leg, fy = yield strength of the

transverse reinforcing, s = transverse reinforcing spacing Al = total area of the longitudinal

reinforcement, fly = yield strength of the longitudinal reinforcing, and Po = perimeter formed

by area Ao.

Here, the interior cross section dimensions of the beam between the centre-lines of the

longitudinal reinforcing are 286mm wide and 582mm deep. The stirrup sets used were XR12

with an leg area of 113mm2 and the longitudinal reinforcement comprised of 3 XD25 bars top

and bottom with a total area of 2946mm2. The stirrup sets were spaced at 150mm and the

yield strengths obtained from material tests for the transverse and longitudinal reinforcing

were 530MPa and 536MPa respectively. This produced a torsional steel contribution, Ts, of

200kNm.

To calculate the torsional twists corresponding to the cracking and ultimate torsion

moments, relationships from Mitchell and Collins (1991) and Park and Paulay (1975) for the

torsional stiffness were used. In this simplified approach, the section is represented by an

Page D-21
equivalent thin walled tube or hollow section with the same overall dimensions. The pre-

cracking stiffness was a function of the section geometry and the post-cracking torsion

stiffness is governed by the deformation of the reinforcing. The pre-cracking stiffness was

calculated using:

4 A02 t
GK G (D.14)
po

Where GK = the pre-cracking torsional stiffness, G = the shear modulus, A0 = area enclosed

by the centreline of the thin walled tube with thickness, t, and po = perimeter of the area. The

equivalent wall thickness can be found by using:

3 Acp
t (D.15)
4 p cp

Where Acp = area enclosed by the outside perimeter of the concrete cross section, pcp =

perimeter of the area Acp. Utilisation of Equations D.14 and D.15 produce the following

result:

3 4 u 196904 3 u (750 u 400)


2
GK 4730 35 u 95531kNm 2
7 1908 4 u 2( 750  400)

A ratio of the post-cracking stiffness to pre-cracking stiffness was calculated using methods

included in Park and Paulay (1975). Using the aspect ratio of the section, and a modified

reinforcing ratio expression, the post-cracking stiffness was found to be 4% of the pre-cracked

stiffness (3821kNm2). It is difficult to accurately assess the post cracking torsional stiffness of

a concrete section, and the effect of flexural plastic hinging during Phase 2 loading would

have had a large effect. The deteriorated state of the concrete within the plastic hinge zone

and cyclic loading with which the reinforcing had been subjected to, means that the post-

cracking stiffness is probably over-estimated.

From the calculated values for the torsional strengths and stiffnesses, the

corresponding twist rotations can be found. The twist rotations were multiplied by the length

Page D-22
of the plastic hinge zone to produce rotations in terms of radians. Considering the torsion

demands and capacity values calculated, Figure D-17 could be generated. The contrast

between the initial stiff pre-cracked response and the post-cracked response can be seen in

Figure D-17. Following Phase 2, the ends of the transverse beams were badly damaged. Thus,

the post-cracked response starts from zero torsional twist in Figure 5-17 and as there is no

concrete contribution due to the state of damage, the ultimate strength is that of the steel only,

as calculated by Equation D.13. The intersection of the torsional demand values gives an idea

as to the torsional twist that would occur, and the trends, more so than the specific values, are

representative of the actual behaviour.

300

250
Torsion Moment (kNm)

200

150

100 Phase I
Phase III
50 Current Experiment Demand
Lindsay (2004) Experiment Demand
0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
Torsional Twist (rads/m)

Figure D-17 Representation of the calculated torsional strengths and demands

D.6 DISPLACEMENT INCOMPATIBILITY

D.6.1 Experimental Results

The relative displacements between the south longitudinal beams and the southernmost

hollow-core unit were measured throughout testing by an array of potentiometers placed

underneath the timber infill topping slab. In addition to this, levelling was undertaken at the

first cycle peak inter-storey drifts during testing, and this was able to give both the relative

displacements, but also a fixed datum to indicate the vertical displacements with relation to

the ground. The following Figures D-18, D-19 and D-20 show the displaced shapes for the

Page D-23
south beams, the first hollow-core unit (Unit 1) and the relative displacement incompatibility

between the floor and beams. A negative displacement incompatibility equates to the flooring

being lower than the beams.


12
Beam 1%
10 Beam -1%
8 Beam 2%
Beam -2%
Vertical Displacement (mm)

6
4
2
0
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 12000
-2
-4
-6
-8
-10
Distance along structure (m m )

(a) Vertical displacement profile of the south beams


6
Slab 1%
Slab -1%
4 Slab 2%
Vertical Displacement (mm)

Slab -2%

0
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 12000

-2

-4

-6
Distance along structure (m m )

(b) Vertical displacement profile of hollow-core unit 1


10
8 1%
-1%
6 2%
Vertical Displacement (mm)

-2%
4
2
0
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 12000
-2
-4
-6
-8
-10
-12
Distance along structure (m m )

(c) Vertical displacement incompatibility between the perimeter beams and floor
Figure D-18 Displacement of the south beams and floor slab during Phase 1

Page D-24
4
Beam 1%
3 Beam -1%
Beam 2%
2 Beam -2%

Vertical Displacement (mm)


Beam 3%
1 Beam -3%
0
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 12000
-1
-2
-3
-4
-5
-6
-7
Distance along structure (m m )

(a) Vertical displacement profile of the south beams


10
Slab 1%
Slab -1%
Slab 2%
5 Slab -2%
Vertical Displacement (mm)

Slab 3%
Slab -3%
0
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 12000

-5

-10

-15

-20
Distance along structure (m m )

(b) Vertical displacement profile of hollow-core unit 1


10
1%
-1%
2%
5 -2%
Vertical Displacement (mm)

3%

0
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 12000

-5

-10

-15
Distance along structure (m m )
(c) Vertical displacement incompatibility between the perimeter beams and floor
Figure D-19 Displacement of the south beams and floor slab during Phase 2

Page D-25
20
Beam 2%
Beam -2%
15 Beam 3%
Beam -3%

Vertical Displacement (mm)


10 Beam 4%
Beam -4%

0
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 12000

-5

-10

-15

-20
Distance along structure (m m )

(a) Vertical displacement profile of the south beams


Distance along structure (m m )
0
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 12000

-5
Vertical Displacement (mm)

-10

-15

Slab 2%
Slab -2%
-20 Slab 3%
Slab -3%
Slab 4%
Slab -4%
-25

(b) Vertical displacement profile of hollow-core unit 1

Distance along structure (m m )


5

0
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 12000
Vertical Displacement (mm)

-5

-10

-15

-20
2%
-25 -2%
3%
-3%
-30 4%
-4%
-35
(c) Vertical displacement incompatibility between the perimeter beams and floor
Figure D-20 Displacement of the south beams and floor slab during Phase 3

Page D-26
The displacement profiles for the beam and floor slab for Phase 1 loading show expected

results. The beams deform in double curvature, and the floor slab in single curvature over the

total span of the building. Figure D-18(c) shows the displacement incompatibility between the

beams and floor, where the maximum relative displacement was 10mm when the flooring was

lower than the frame beams. Reasonably symmetric behaviour was observed between the

positive and negative cycles, and the maximum displacement incompatibility was seen to

occur around 1m either side of the central column (6100mm). Similarly, the results from

Phase 2 (Figure D-19) were expected. The displacement incompatibility relationship shows

that the flooring was a maximum of 5mm higher than the beams during negative drifts when

the transverse beams at the south end were inclining upwards, and during the positive drift

cycles, the flooring was a maximum of 14mm lower than the frame beams.

The results for Phase 3 as seen in Figure D-20 show interesting results. The deformed

shape of the beams at various drift amplitudes shows an expected curvature pattern, but the

displacement of the floors shows a distinct drop of over 16mm. During the experiment, it was

observed that the transverse beams began to drop during Phase 2 loading. Most of the rotation

within the plastic hinges at the end of the transverse beams was centred at the beam-column

interfaces where a significant crack had formed. During the later stages of Phase 2, it was

noticed that the beams were dropping and a vertical dislocation was occurring at these large

end cracks. This is displayed by the vertical displacement results whereupon comparing Phase

1 and 3 (Figure 5-18(b) and 5-20(b)) the amount of the dropping of the floor units is noticed.

Despite this, and as Figure 5-20(c) shows, the same trends of displacement incompatibility

were seen, but the vertical dropping meant that the floor remained predominantly lower than

the southern beams. At the floorings units highest point, the floor was 1mm higher than the

beams, and at its lowest, 33mm below the beams. The maximum incompatibility results were

seen to occur at approximately 1m either side of the central column.

Page D-27
D.6.2 Analytical Modelling

The analysis involved the construction of an analytical model in the inelastic time history

program Ruaumoko (Carr, 2003), utilising a plasticity concentrated approach. A two-

dimensional model was been set up to model the longitudinal lateral loading of the front

frame, and consists of members representing the frame and floor elements and springs

connecting the floor to the frame. Different spring members were used to model the level of

fixity of the end support conditions, and along the frame between the frame and flooring.

Sensitivity analyses were performed examining the effect of various parameters; different end

support conditions, different lateral connection details, different bay lengths and numbers of

bays and different flooring types. The effects of the different parameters were seen by

examining both the displacement incompatibility patterns as well as the overall system base

shear capacities.

From the sensitivity analyses, it was seen that the end connection detail, through the

stiffness of the end springs, had a large influence on system strength capacity and

displacement incompatibility values. This is shown by Figure D-21, and is matched to a

degree by the higher ultimate base shear capacity of the current experimental structure than

the equivalent Lindsay (2004) base shear capacity.

3500

3000

FIXED k=9.9x108
2500
k=1x108
Base shear (kN)

2000

k=1x107
1500
PINNED k=0,1,1000,100000

1000

500

0
0.0 1.0 2.0 3.0
Inclination (%)

Figure D-21 Modelling the effect of the end connection detail on the base shear capacity

Page D-28
Modifying the side spring characteristics was done by altering the spring stiffnesses or the

yield point representing the point at which cracking or tearing would occur. Adjusting the

stiffness was seen to have a large effect on the displacement compatibility trends, with the

obvious increased deformations for softer springs. The more stiff the springs also

corresponded to an increase in the strength capacity of the system due to the greater ability to

transfer the floor inertia forces to the frame. Figure D-22 shows the effect on the base shear

capacity of the structure for various side connection detail stiffness values.

3500

RIGID k=9.9x108
3000 8
k=1x10
k=1x107
2500
k=0,1,1000 &100000
Base shear (kN)

"SOFT"
2000

1500

1000

500

0
0.0 1.0 2.0 3.0
Inclination (%)

Figure D-22 Modelling the effect of the side connection detail on the base shear capacity

Different bay lengths, using 4m, 6m and 8m showed no major effect on the displacement

trends. Altering the floor system types for 200, 300 and 400 series hollow-core was done by

changing the member stiffness to 0.35EI, 1EI and 2.2EI respectively. This was seen to have a

significant effect with the most severe displacement incompatibility values observed for the

stiff 400 series floor. This can be explained by the fact that the more flexible 200 and 300

series flooring is able to deform with the beams in a double curvature manner (within the

model, but not in reality), whereas the stiff nature of the 400 series means the floor deforms in

a single curvature way.

Page D-29
Figure D-23 shows the displacement incompatibility results from the experimental tests and

modelling for both the Lindsay (2004) and current experiments. The lateral spring properties

modelling the infill connection were kept the same, but the end spring properties were altered

to model the pinned type seating connection for Lindsay (2004) and the rigid connection in

the present experiment. The results show promising resemblance between the experiment and

analytical model. The overall deformation trends are matched although the model slightly

underestimates the amount of incompatibility that occurs. For the rigid end conditions, the

maximum displacement incompatibility occurred on both sides adjacent to the central column,

as the floor unit deformed in single curvature due to the fixity at the ends. For the pinned end

conditions, which allowed the floor to simply sag, displacement incompatibility peaks were

seen close to the supports as well as either side of the central column, and the displacement

incompatibility values were generally higher than for the fixed end case.
10
8
6
4
Displacement (mm)

2 Position along fram e (m m )

0
-2 0 2000 4000 6000 8000 10000 12000

-4
-6 Experiment 1%
-8 Model 1%
Experiment 2%
-10 Model 2%

-12

(a) Observed and modelled displacement incompatibility from Lindsay (2004)


10
8
6
4
Displacement (mm)

2
Position along fram e (m m )
0
-2 0 2000 4000 6000 8000 10000 12000

-4
-6 Experiment 1%
-8 Model 1%
Experiment 2%
-10 Model 2%

-12

(b) Observed and modelled displacement incompatibility from the current experiment
Figure D-23 Observed and modelled displacement incompatibility

Page D-30
D.7 BEAM ELONGATION

The phenomenon of beam elongation is explained in Sections 2 and 3 of this thesis. Matthews

et al (2004) developed a rainflow-counting method to predict the amount of elongation that

occurs. The assumption of the eccentricity values, ecr, is a key part of the theoretical

predictions, and the same values used by Matthews (2004) and later by Lindsay (2004) were

used in these calculations. The following Figures D-24, D-25 and D-26 plot the observed

beam elongation results as well as the predicted response for the three loading phases of the

experiment.

14 14
12 Elongation (mm) Experiment 12
Elongation (mm)

10 Theo ry 10
8 8
6 7 6
Experiment 4 4
Theo ry 2 2
0 0
-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3
Int e r- S t o re y D rif t ( %) Inter-Storey Drift (%)

(a) South west plastic hinge elongation (b) South centre west hinge elongation
14 20
18
12
Elongation (mm)

Elongation (mm)

16
10 14
12
8
10
7 6 8
Experiment 6
4
4
Experiment Theo ry
2 2
Theo ry
0
0
-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3
Int e r- S t o re y D rif t ( %) Int e r- S t o re y D rif t ( %)

(c) South centre east hinge elongation (d) South east hinge elongation
45
40
Elongation (mm)

35
30
25
20
15
Experiment 10
Theory 5
0
-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5
Inter-Storey Drift (%)

(e) Total frame elongation


Figure D-24 Beam elongation results for Phase 1
Page D-31
20 20
18 18

Elongation (mm)
Elongation (mm) 16 16
14 14
12 12
10 10
8 8
6 6
Experiment 4 Experiment 4
Theo ry 2 Theo ry 2
0 0
-4 -3 -2 -1 0 1 2 3 4 -4 -3 -2 -1 0 1 2 3 4
Inter-Storey Drift (%) Int e r- S t o re y D rif t ( %)

(a) South west plastic hinge elongation (b) North west hinge elongation
20 20
18 18
Elongation (mm)

Elongation (mm)
16 16
14 14
12 12
10 10
8 8
6 6
Experiment 4 Experiment 4
Theo ry 2 Theo ry 2
0 0
-4 -3 -2 -1 0 1 2 3 4 -4 -3 -2 -1 0 1 2 3 4
Int e r- S t o re y D rif t ( %) Inter-Storey Drift (%)

(c) South east hinge elongation (d) North east hinge elongation
Elongation

35
(mm)

30

25

20

15

10
Phase II East Bay
5 Phase II West Bay
Theory
0
-4 -3 -2 -1 0 1 2 3 4
Inter-storey Drift (%)

(e) Total elongation (east and west bays)


Figure D-25 Beam elongation results for Phase 2
120
Elongation
(mm)

100

80

60

40

Phase I
20 Phase III
Theory
0
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
Inter-storey Drift (%)

Figure D-26 Total Beam elongation for Phase 1 and 3

Page D-32
Theoretical Predictions: Longitudinal Loading

Yield Drift
0.5 %
0.005 rad
ecr: External Hinges
Distance (mm) ecr*Ty
0.425D = 318.75 1.59375
0.475D = 356.25 1.78125
ecr: Internal Hinges
Distance (mm) ecr*Ty
0.225D = 168.75 0.84375
0.275D = 206.25 1.03125

Drift Plastic Hinge of South Frame Bay Total Elongation


(%) (rads) SE SCE SCW SW East West (mm)
0 0 0 0 0 0 0 0 0
0.5 0.005 1.59375 1.03125 0.84375 1.78125 2.625 2.625 5.25
0 0 0 0 0 0 0 0 0
-0.5 -0.005 1.78125 0.84375 1.03125 1.59375 2.625 2.625 5.25
0 0 0 0 0 0 0 0 0
1 0.01 3.1875 2.0625 1.6875 3.5625 5.25 5.25 10.5
0 0 1.59375 1.03125 0.84375 1.78125 2.625 2.625 5.25
-1 -0.01 4.96875 2.90625 2.71875 5.15625 7.875 7.875 15.75
0 0 3.1875 2.0625 1.6875 3.5625 5.25 5.25 10.5
1 0.01 4.78125 3.09375 2.53125 5.34375 7.875 7.875 15.75
2 0.02 7.96875 5.15625 4.21875 8.90625 13.125 13.125 26.25
0 0 6.375 4.125 3.375 7.125 10.5 10.5 21
-1 -0.01 8.15625 4.96875 4.40625 8.71875 13.125 13.125 26.25
-2 -0.02 11.71875 6.65625 6.46875 11.90625 18.375 18.375 36.75
0 0 9.9375 5.8125 5.4375 10.3125 15.75 15.75 31.5
2 0.02 11.53125 6.84375 6.28125 12.09375 18.375 18.375 36.75
3 0.03 14.71875 8.90625 7.96875 15.65625 23.625 23.625 47.25
0 0 13.125 7.875 7.125 13.875 21 21 42
-2 -0.02 14.90625 8.71875 8.15625 15.46875 23.625 23.625 47.25
-3 -0.03 18.46875 10.40625 10.21875 18.65625 28.875 28.875 57.75
0 0 16.6875 9.5625 9.1875 17.0625 26.25 26.25 52.5
3 0.03 18.28125 10.59375 10.03125 18.84375 28.875 28.875 57.75
4 0.04 21.46875 12.65625 11.71875 22.40625 34.125 34.125 68.25
0 0 19.875 11.625 10.875 20.625 31.5 31.5 63
-3 -0.03 21.65625 12.46875 11.90625 22.21875 34.125 34.125 68.25
-4 -0.04 25.21875 14.15625 13.96875 25.40625 39.375 39.375 78.75
0 0 23.4375 13.3125 12.9375 23.8125 36.75 36.75 73.5
4 0.04 25.03125 14.34375 13.78125 25.59375 39.375 39.375 78.75
5 0.05 28.21875 16.40625 15.46875 29.15625 44.625 44.625 89.25
0 0 26.625 15.375 14.625 27.375 42 42 84
-4 -0.04 28.40625 16.21875 15.65625 28.96875 44.625 44.625 89.25
-5 -0.05 31.96875 17.90625 17.71875 32.15625 49.875 49.875 99.75
0 0 30.1875 17.0625 16.6875 30.5625 47.25 47.25 94.5

Page D-33
Theoretical Predictions: Transverse Loading

Yield Drift
0.5 %
0.005 rad
ecr: External Hinges
Distance (mm) ecr*Ty
0.425D = 318.75 1.59375
0.475D = 356.25 1.78125

Drift Plastic Hinges East/West Bay Total


(%) (rads) North South (mm)
0 0 0 0 0
0.5 0.005 1.59375 1.78125 3.375
0 0 0 0 0
-0.5 -0.005 1.78125 1.59375 3.375
0 0 0 0 0
1 0.01 3.1875 3.5625 6.75
0 0 1.59375 1.78125 3.375
-1 -0.01 4.96875 5.15625 10.125
0 0 3.1875 3.5625 6.75
1 0.01 4.78125 5.34375 10.125
2 0.02 7.96875 8.90625 16.875
0 0 6.375 7.125 13.5
-1 -0.01 8.15625 8.71875 16.875
-2 -0.02 11.71875 11.90625 23.625
0 0 9.9375 10.3125 20.25
2 0.02 11.53125 12.09375 23.625
3 0.03 14.71875 15.65625 30.375
0 0 13.125 13.875 27
-2 -0.02 14.90625 15.46875 30.375
-3 -0.03 18.46875 18.65625 37.125
0 0 16.6875 17.0625 33.75

D.8 LOW CYCLE FATIGUE FRACTURE OF REINFORCEMENT

A low cycle fatigue theory proposed by Dutta and Mander (2001) enables the plastic rotation

life capacity of reinforcement to be calculated. The theory has been outlined in Section 2 and

Appendix A to predict the fracture of the reinforcement in the sub-assemblage experiments,

and Section 4 to assess the fracture of longitudinal reinforcement of the main beams during

the super-assemblage experiment. The following calculations outline the derivation of the

theoretical fatigue limits of the reinforcement, for longitudinal loading.

Page D-34
The equivalent number of equi-amplitude cycles to the reference drift (5%) experienced

during the experiment is calculated as the following:

Tref= 5
Nmax= 5.12
q ncycles (T/Tref)2
0.5 3 0.03
1 2 0.08
2 3 0.48
3 2 0.72
4 2 1.28
5 2 1.9
6 4.49

This calculation shows that the reinforcement withstood 4.5 equivalent cycles (NF-Expt) to 5%

during the experiment before fracture occurred. By re-arranging the low cycle fatigue

equation from Dutta and Mander (2001), the theoretical equivalent number of 5% cycles,

NF-Theory, that the reinforcement will be able to sustain prior to fracture can be found from the

following:

2
D L p 1 L p
0.16 1 
1 D' D 2 L
N F Theory (D-16)
2 T T y

2
750 511 1 511
0.16 1 
N F Theory
1
636 750 2 2675 3.6
2 0.05  0.0064

Therefore, the theoretical number of cycles at an amplitude of 5% required to cause low cycle

fatigue fracture of the reinforcement is 3.6.

Page D-35
D.9 OBSERVED BOWSTRING EFFECT

To account for the presence of the floor slab in calculating the overall strength of the

structure, simple calculations were used to quantify the bowstring effect. Beam elongation is

restrained due to the presence of the floor slab, and this restraint induces large compression

forces within the longitudinal beams. To calculate this compression force, a simplified

approach is used which calculates the eccentricity or depth of the compression arch. From

this, the magnitude of the tension force can be found which is a function of the capacity of the

reinforcement of the hollow-core end connections, the friction of the hollow-core seating and

the amount of diaphragm steel orthogonal to the tension tie. The calculation is as follows:

wl 2 Pl
Te  (D.16)
8 4

where e = midspan eccentricity of the tie force with respect to the centroid of the compression

strut, T = combined tie force, l = overall span of arch, P = the tension force provided by the

drag bar reinforcement and w = uniformly distributed load given by the diaphragm reinforcing

orthogonal to the tension tie. As the tie force, T, is a function of the tension reinforcement

capacity at the ends of the floor, it can be represented as:

2eAs f y
T  2 PWle (D.17)
s

where As = end connection reinforcing area (starter bars and hairpin bars), fy = end connection

reinforcing yield strength, s = end connection reinforcing spacing, P = coefficient of friction

for the bearing strip and W = weight of the hollow-core. The uniformly distributed load given

by the diaphragm reinforcing orthogonal to the tension tie, w, can be found by:

Asd f yd
w (D.18)
s sd

where Asd = area of the diaphragm steel, fyd = yield strength of the diaphragm reinforcing and

ssd = spacing of the diaphragm reinforcing.

Page D-36
Using this formulation, and the measured material properties of the reinforcement, an

eccentricity of 2.5m was found. This matches well with the cracking pattern witnessed

throughout testing. The eccentricity value allowed a tension force of 1254kN to be calculated.

By equilibrium, this tension force is resisted by a compression force of the same magnitude,

and this compression force has a significant effect on the capacity of the system and notably

the strength of the beam hinges joining the central column.

The compression axial force was taken to act at the level of the topping slab and

further simplified to occur at the top extreme fibre. Therefore, the moment capacity of both of

the hinges was found to be:

M beams ( As  As ) f y (d  d ' )  Nd (D-19)


+ -
In which As = area of tension reinforcement; As = area of compression reinforcement; fy =

yield strength of tension reinforcement; d = the distance from the extreme compression fibre

to the centroid of the tension reinforcement; d = the distance from the extreme compression

fibre to the centre of the compression block; and N = the axial compression force due to the

bowstring effect. Using this formulation, the moment demand on the interior column, as

found by the sum of the two hinges framing into the south centre column, was found to be

1859kNm. This is equivalent to the nominal moment capacity (1004kNm) when the

bowstring effect is omitted.

Page D-37
D.10 SUPER-ASSEMBLAGE CAPACITY MECHANISM CALCULATIONS

The capacity mechanisms used to calculate the theoretical pushover curves for both the

longitudinal and transverse loading directions, is outlined in this sub-section. The theoretical

curve is defined by two points: the onset of yielding and the point of full plastification.

Yielding was assumed to occur at 0.64% as calculated in Section D.1. Plastification was

assumed to occur at 2% in keeping with the work done by Matthews (2004) and the

observed crack widths during this experiment. The contributions to the overall strength are

from the nominal moment strengths of the beams and for negative moment capacities,

additional strength is derived from activation of the starter bars or end connection

reinforcement and diaphragm reinforcement. The bowstring effect, described in section D.8 is

also recognised.

D.10.1 Phase 1 and 3: Longitudinal

For the external hinges of the structure, in this case for longitudinal loading, the southeast and

southwest hinges, the negative moment capacity is enhanced by the activation of starter bars

from the transverse beams. These starter bars are progressively activated with increasing

lateral drift. For the yield point, the contributing slab width was taken as the width of the infill

topping slab. At the point of full plastification, the width of activated slab was taken to be

3.05m, or approximately half the span of the transverse beams. At 2% drift the width of the

crack between the beam and slab over the full length of the transverse beam, was over 5mm

signifying that the bars were yielding and contributing to the strength of the system. The

moment capacity values calculated were extrapolated to the column centreline in order to

calculate the overall system base shear, and for this, the plastic hinges were assumed to form

at 0.5D (375mm) from the column face.

Page D-38
Initial Yield

Southeast and Southwest hinges:

M+ = 502kNm (at column centreline) 663kNm

M- = 502kNm + 3 infill starters = 570kNm (at column centreline) 752kNm

South central hinges: (accounting for axial compression force due to bowstring effect)

M+ = 502kNm (at column centreline) 663kNm

M- = 1357kNm (at column centreline) 1791kNm

Full Plastification

Southeast and Southwest hinges:

M+ = 502kNm (at column centreline) 663kNm

M- = 502kNm + 11 infill starters = 707kNm (at column centreline) 933kNm


(0.8jd used to allow for the hairpin reinforcement in compression reducing the internal lever arm)

South central hinges:

M+ = 502kNm (at column centreline) 663kNm

M- = 1357kNm + 3 diaphragm bars = 1425kNm (at column centreline) 1881kNm

Also, as the base shear capacity of the entire building is being predicted, the negative moment

contribution from the top bars at the northern ends of the transverse beams need to be taken

into account.

Overall Summary and Base Shear Calculation:

Hinge: NW SW SC SE NE 6M Vb
M
hc

Yield - 663 2453 752 - 3868 1105kN


Positive
Drift Plastification - 663 2543 953 191 4350 1243kN

Yield - -752 -2453 -663 - -3868 -1105kN


Negative
Drift Plastification -191 -953 -2543 -663 - -4350 -1243kN

Page D-39
As the experimental results were zeroed at the commencement of testing, the experimental

results do not show the residual or natural moments of the structure. During construction, all

of the beams were propped during the placement of the floors and the casting of the topping

concrete. Upon removal of these supports, the eccentric weight of the floor would have

caused the east and west columns to incline inwards (tops of the columns move towards the

centre of the building, and the bases of the columns move outwards), and thus cause positive

moments within the southeast and southwest hinge zones. This artificially shifted the

hystereses off the true zero moment. By way of a simple elastic analysis using the gross

section properties of the structure, this initial moment was calculated as 120kNm, and the

hysteretic plots for both the southeast and southwest hinges in the longitudinal direction have

been shifted by this amount.

Figure D-27 shows the comparisons between the predicted and actual strengths. An

upper bound solution was also provided based on 25% overstrength.

D.10.2 Phase 2: Transverse

A similar method of analysis was used to calculate the capacity of the super-assemblage in the

transverse direction. In this case, the strengths of the transverse beam hinges and the starter

bars from the south longitudinal beams were the contributing sources of strength. Due to the

presence of the steel tie frame along the northern edge of the structure, no starter bars were

present to enhance the negative moment capacity of the northern hinges, and hence slightly

non-symmetrical behaviour is present. The amount of starter bars activated within the

longitudinal beams was taken to be the same as in the longitudinal loading direction; at initial

yield, 1m was the contributing slab width, and at full plastification, the slab width was half

the bay width, or 3m.

Page D-40
Initial Yield

Southeast and Southwest hinges:

M+ = 466kNm (at column centreline) 531kNm

M- = 466kNm + 3 infill starters = 532kNm (at column centreline) 607kNm

Northeast and Northwest hinges:

M+ = 466kNm (at column centreline) 531kNm

M- = 466kNm (at column centreline) 531kNm

Full Plastification

Southeast and Southwest hinges:

M+ = 466kNm (at column centreline) 531kNm

M- = 466kNm + 9 infill starters = 663kNm (at column centreline) 756kNm

Northeast and Northwest hinges:

M+ = 466kNm (at column centreline) 531kNm

M- = 466kNm (at column centreline) 531kNm

Overall Summary and Base Shear Calculation:

Hinge: NW SW SE NE 6M Vb
M
hc

Yield 531 531 531 531 2124 607kN


Positive
Drift Plastification 531 615 531 531 2124 607kN

Yield -531 -607 -607 -531 -2276 -650kN


Negative
Drift Plastification -531 -756 -756 -531 -2968 -848kN

These strength predictions for Phase 2 along with Phase 1 and 3 longitudinal loading are

shown in the following Figures. Section 4 of this thesis discusses the results and explains the

comparisons between the actual and theoretical strengths, whether the comparisons are good

or otherwise.

Page D-41
1800
1600

Base Shear
Force (kN)
1400
1200
1000
800
600
400
200
Inter-Storey Drift (%)
0
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
-200
-400
-600
-800
Experiment Phase 3
-1000
Experiment Phase 1
-1200
Nominal Prediction
-1400 25% Overstrength Prediction
-1600 40% Overstrength Prediction
-1800

(a) Phase 1 and 3 Longitudinal loading: Experimental base shear capacity and theoretical
pushover curve.
1200
Base Shear
Force (kN)

900

600

300

Inter-Storey Drift (%)


0
-4 -3 -2 -1 0 1 2 3 4

-300

-600 Experiment (pre-torsion)

Experiment (post-torsion)

-900 Nominal Prediction

25% Overstrength Prediction

-1200

(b) Phase 2 Transverse loading: Experimental base shear capacity and theoretical pushover
curve.
Figure D-27 Base shear capacity: experimental and theoretical relationships.

Page D-42
1200

Moment (kNm)
1000
800

600
400
200
R o t a t io n ( %)
0
-6 -5 -4 -3 -2 -1 -200 0 1 2 3 4 5 6

-400
Phase 1
-600
Phase 3
-800 Nominal Strength
-1000 25% Overstrength
-1200

(a) Moment rotation behaviour for the southwest hinge of the south frame
Moment (kNm)

3000
2500
2000
1500
1000
500
R o t a t io n ( %)
0
-6 -5 -4 -3 -2 -1 -500 0 1 2 3 4 5 6

-1000
Phase 1
-1500
Phase 3
-2000 Nominal Strength
-2500 25% Overstrength
-3000

(b) Moment rotation behaviour for the two centre hinges of the south frame framing into the
south centre column
Moment (kNm)

1200
1000
800

600
400
200
R o t a t io n ( %)
0
-6 -5 -4 -3 -2 -1 -200 0 1 2 3 4 5 6

-400 Phase 1
-600 Phase 3
-800 Nominal Strength
-1000 25% Overstrength
-1200

(c) Moment rotation behaviour for the southeast hinge of the south frame
Figure D-28 Phase 1 and 3 Moment rotation behaviour: experimental and theoretical
relationships.

Page D-43
800

Moment (kNm)
600

400

200

Rotation (%)
0
-4 -3 -2 -1 0 1 2 3 4
-200

Pre-Torsion
-400
Post-Torsion
-600 Nominal Strength
25% Overstrength
-800

(a) Moment rotation behaviour for the north hinge of the west frame
1200
Moment (kNm)

1000
800
600
400
200
Rotation (%)
0
-4 -3 -2 -1 -200 0 1 2 3 4

-400
-600 Pre-Torsion
Post-Torsion
-800
Nominal Strength
-1000 25% Overstrength
-1200

(b) Moment rotation behaviour for the south hinge of the west frame
Figure D-30 Phase 2 Moment rotation behaviour: experimental and theoretical relationships
for the west frame.

Page D-44
800

Moment (kNm)
600

400

200

Rotation (%)
0
-4 -3 -2 -1 0 1 2 3 4
-200

-400 Pre-Torsion
Post-Torsion
-600 Nominal Strength
25% Overstrength
-800

(a) Moment rotation behaviour for the north hinge of the east frame
1000
Moment (kNm)

800

600

400

200
Rotation (%) 0
-4 -3 -2 -1 0 1 2 3 4
-200

-400 Pre-Torsion
Post-Torsion
-600
Nominal Strength
-800 25% Overstrength

-1000

(b) Moment rotation behaviour for the south hinge of the east frame

Figure D-30 Phase 2 Moment rotation behaviour: experimental and theoretical


relationships for the east frame.

Page D-45
D.11 REFERENCES

Carr A.J, 2003, Ruaumoko 2D User Manual, Computer Program Library, Department of Civil

Engineering, University of Canterbury, Christchurch, New Zealand.

Dutta A and Mander J.B, 2001, Energy Based Methodology for Ductile Design of Concrete

Columns, Journal of Structural Engineering, Vol. 127(12), December, pp 1374-1381.

Lindsay R.A, 2004, Experiments on the seismic performance of hollow-core floor systems in

precast concrete building, Masters Thesis, Department of Civil Engineering,

University of Canterbury, Christchurch, New Zealand.

Matthews J.G, 2004, Hollow-core floor slab performance following a severe earthquake, PhD

Thesis, Department of Civil Engineering, University of Canterbury, Christchurch,

New Zealand.

Matthews J.G, Mander J.B, Bull D.K, 2004, Prediction of beam elongation in structural

concrete members using a rainflow method, 2004 NZSEE Conference, Rotorua, NZ,

Paper 27, 8 pp.

Mitchell M.P and Collins D, 1991, Prestressed Concrete Structures, Prentice Hall,

Englewood Cliffs, N.J, USA.

NZS3101, 1995, Concrete Structures Standard, NZS3101, Parts 1 & 2, Standards New

Zealand, Wellington, New Zealand.

NZS3101, 2003, Draft Amendment No. 3 to 1995 Standard (NZS3101), Standards New

Zealand, Wellington, New Zealand.

Park R and Paulay T, 1975, Reinforced Concrete Structures, J. Wiley and Sons Inc, New

York.

Priestley M.J.N and Kowalsky M.J, 2000, Direct Displacement Based Design of Concrete

Buildings, Bulletin of the New Zealand Society for Earthquake Engineering,

December, Vol. 33, No.4, pp 421-444.

Page D-46

Das könnte Ihnen auch gefallen