Você está na página 1de 228

CUREE Publication No.

CKIV-02

SEISMIC PERFORMANCE ASSESSMENT OF


FLAT PLATE FLOOR SYSTEMS

Thomas H.-K. Kang


Changsoon Rha
John W. Wallace
University of California, Los Angeles

Katsuya Igarashi
Norio Suzuki
Kajima Corporation

CUREE

CUREE-Kajima Joint Research Program


Phase IV

April 2003
This page left intentionally blank.
CUREE Publication No. CKIV-02

SEISMIC PERFORMANCE ASSESSMENT OF


FLAT PLATE FLOOR SYSTEMS

Thomas H.-K. Kang


Changsoon Rha
John W. Wallace
University of California, Los Angeles

Katsuya Igarashi
Norio Suzuki
Kajima Corporation

April 2003

CUREE
Consortium of Universities for Research in Earthquake Engineering
1301 S. 46th St.
Richmond, CA 94804-4698
tel.: 510-231-9557 fax: 510-231-5664
e-mail: curee@curee.org website: www.curee.org
This page left intentionally blank.
Abstract

An investigation into the lateral load response of reinforced concrete flat plate frames utilizing
stud rails for shear reinforcement at the slab - column connections was carried out. The two by
two bay, two story specimens were approximately one-third scale representations of typical slab -
column frames constructed in moderate-to-high seismic zones in the United States. One of the
specimens consisted of a conventional reinforced concrete flat plate (RC specimen), whereas the
other specimen consisted of nominally reinforced flat plate with post-tensioning reinforcement
(PT specimen). The specimens were subjected to gravity loads and increasing intensity of
uniaxial base acceleration histories on the shake table at the Earthquake Engineering Research
Center at UC Berkeley’s Richmond Field Station. During testing, data were collected from 193
channels and five or six video cameras to assist in assessing the behavior of the specimens.

Although slab-column punching failures occurred during the tests, lateral drift ratios of 3% and
4% were achieved for the RC and PT frames, respectively, with relatively little loss of lateral
load capacity. Analytical models including column cracking, and based on using an effective slab
width model with an effective width factor α of 0.8 and 0.65 and a cracking factor β of 1/3 and
1/2 for the RC and PT specimens, respectively, resulted in good correspondence between
experimental and experimental responses for low-to-moderate levels of shaking. Base shear
versus top level displacement relations for the analytical models also captured the nonlinear
envelop responses for the more intense shaking levels reasonably well.

Overall, the results indicate that the slab - column frames can be designed to have sufficient drift
capacity to be used as a non-participating frame, or as a primary lateral force resisting system in
low-to-moderate seismic regions.

iii
Acknowledgements

This research was conducted with funding provided by Phase IV of the joint CUREE – Kajima
Research Program. The authors wish to acknowledge the dedication of the Joint Oversight
Committee for the Phase IV CUREE – Kajima research program. The active participation of
Kajima Corporation researchers Dr. Katsuya Igarashi and Dr. Norio Suzuki throughout the
project, as well as their active participation and interest in the research program are gratefully
acknowledged.

Don Clyde and Wesley Neighbour, from the UC Berkeley EERC/PEER Richmond Field Station,
are thanked for providing valuable advice during the construction, instrumentation, and testing of
the specimens. John Schuller at RPS Inc is thanked for his helpful suggestions during the post-
tensioning procedure. Professors Jack P. Moehle and Stephan A. Mahin at UC Berkeley are
thanked for their support and advice throughout the testing phase of the research.

Special thanks are due to Dr. Shahkzod Takhirov for his assistance with the rebar strain gauges,
baseplates, and grouting, as well as for his valuable advice. The authors also would like to
acknowledge UC Berkeley students Carlos Machado, Aditya Hariharan, Joong Hwan Kim, Yoon
Bong Shin, Karthik Sethuraman, Natalia Carse, who assisted the instrumentation, and David
Maclam, the machinist at EERC/PEER, for his excellent work. Senior Design Engineer Harold
Kasper and PhD student Murat Melek, both of the Department of Civil & Environmental
Engineering at UCLA, are thanked for their assistance with materials testing.

The studrails® used in the test program was supplied courtesy of Decon® USA Inc.

iv
Table of Contents
Page

Abstract......................................................................................................................................... iii
Acknowledgements ...................................................................................................................... iv
Table of Contents ...........................................................................................................................v
List of Tables .............................................................................................................................. viii
List of Figures............................................................................................................................... ix

Chapter 1 – Introduction...............................................................................................................1
1.1 Introduction....................................................................................................................1
1.2 Summary ........................................................................................................................4
1.3 Report Organization.......................................................................................................5

Chapter 2 – Literature Review .....................................................................................................7


2.1 Lateral-load Stiffness .....................................................................................................7
2.2 Lateral-load Strength....................................................................................................12
2.2.1 Yield Line Failure ...............................................................................................12
2.2.2 Eccentric Shear Stress Model .............................................................................13
2.3 Lateral-load Ductility...................................................................................................18
2.4 Post-punching Behavior...............................................................................................19
2.5 Post-tensioned Flat Plate..............................................................................................20
2.6 Shake Table Tests of Flat Plate Systems .....................................................................23
2.7 Summary ......................................................................................................................26

Chapter 3 – Specimen Design and Construction ......................................................................29


3.1 RC Specimen ..............................................................................................................29
3.2 PT Specimen ................................................................................................................34

Chapter 4 – Instrumentation and Experimentation .................................................................39


4.1 Test Setup.....................................................................................................................39
4.2 Instrumentation ............................................................................................................43
4.2.1 Table Instrumentation .........................................................................................44
4.2.2 Strain Gauges ......................................................................................................44

v
4.2.3 Load Cells ...........................................................................................................45
4.2.4 Accelerometers....................................................................................................46
4.2.5 Displacement Gauges..........................................................................................47
4.3 Testing..........................................................................................................................50

Chapter 5 – Experimental Results..............................................................................................53


5.1 Base Shear versus Top Relative Displacement............................................................53
5.2 Column Moment Evaluation........................................................................................55
5.3 Slab Moment Evaluation..............................................................................................58
5.4 Observed Damage ........................................................................................................60
5.5 Punching Failure ..........................................................................................................61
5.6 Drift Capacity of Slab-column Connections at Punching............................................63
5.7 Monitoring and Variation of Post-tensioning Forces...................................................68

Chapter 6 – Analytical Studies ...................................................................................................73


6.1 Introduction..................................................................................................................73
6.2 Description of Analytical Models ................................................................................73
6.3 Analytical Models: Low-to-moderate Level Responses ..............................................77
6.3.1 Period Comparisons ............................................................................................78
6.3.2 Base Shear and Top Relative Displacement Comparisons .................................80
6.4 Analytical Models: Nonlinear (Yielding) Responses ..................................................81
6.4.1 Punching Failure prior to Slab Flexural Yielding...............................................82
6.4.2 Punching Failure after Slab Flexural Yielding within c + 3h .............................83
6.4.3 Punching Failure after Slab Flexural Yielding within Column Strip..................85
6.4.4 Response of Static Push-over Analyses ..............................................................86

Chapter 7 – Summary and Conclusions ....................................................................................87


7.1 Summary ......................................................................................................................87
7.2 Conclusions..................................................................................................................87

vi
References .....................................................................................................................................93

Tables ............................................................................................................................................99

Figures.........................................................................................................................................127

Appendix A.................................................................................................................................199

vii
List of Tables
Page

Table 2-1 Ratios of measured and calculated shear strength using eccentric
shear stress model .......................................................................................................99
Table 2-2 Test variables and failure models (Dilger and Shatila, 1989) ...................................100
Table 2-3 Test and response summary (Moehle and Diebold, 1984) ........................................100
Table 3-1 Concrete compressive strength results of core samples ............................................101
Table 3-2 Reinforcing steel ratios for various conditions..........................................................101
Table 3-3 Concrete compressive strength results of cylinders...................................................102
Table 3-4 Material properties of steel bars and seven-wire strands...........................................103
Table 4-1 Channel list of RC specimen .....................................................................................104
Table 4-2 Channel list of PT specimen......................................................................................110
Table 4-3 Characteristic of original ground motion...................................................................115
Table 4-4 Free vibration test results (RC specimen)..................................................................116
Table 4-5 Free vibration test results (PT specimens).................................................................116
Table 5-1 Various dimensions of displacement gauges (See Fig. 5-9)......................................117
Table 5-2 Drift capacity at punching of the RC specimen.........................................................120
Table 5-3 Drift capacity at punching of the PT specimen..........................................................121
Table 5-4 Post-tensioning ..........................................................................................................122
Table 5-5 Post-tensioning force change during testing..............................................................123
Table 6-1 Period comparisons for elastic element model ..........................................................124
Table 6-2 Moment capacities of springs for each specimen ......................................................126

viii
List of Figures
Page

Fig. 1-1 Slab – column connection with stud-rails (RC specimen) ...........................................127
Fig. 1-2 Slab – column critical section and shear stress distributions .......................................127
Fig. 1-3 Relationships between drift and gravity shear from experiments ................................128
Fig. 1-4 Punched connection with stud-rails superimposed on failure plan ..............................128
Fig. 1-5 Test specimens .............................................................................................................129
Fig. 2-1 Stiffness models for slab – column frame ....................................................................130
Fig. 2-2 Effective beam width model proposed by Luo et al. (1994; 1995) ..............................130
Fig. 2-3 Slab flexural yielding across the full width..................................................................131
Fig. 2-4 Anchored flexural reinforcement within the transfer width .........................................131
Fig. 2-5 Shear and moment transfer assumed in eccentric shear stress model ..........................132
Fig. 2-6 Critical sections outside of shear-reinforced zone........................................................132
Fig. 2-7 Ratio between measured and calculated moment transfer strengths using the effective
transfer width, c1 + 2c2 (Moehle, 1988) ......................................................................133
Fig. 2-8 Effect of gravity load on drift .......................................................................................133
Fig. 2-9 Tests of isolated specimens conducted by Trongtham and Hawkins (1977) ...............134
Fig. 2-10 Comparison of lateral load-deflection relationships
(Hawkins et al., 1975; 1975; 1977) ............................................................................134
Fig. 2-11 Moment-deflection relationships for four specimens (Foutch et al., 1990) ...............135
Fig. 2-12 Load vs. displacement envelopes (Qaisrani, 1993) ....................................................135
Fig. 2-13 Test structure on shake table (Moehle and Diebold, 1984)........................................136
Fig. 2-14 Envelope relationship between top displacement and base shear
(Moehle and Diebold, 1984) ......................................................................................136
Fig. 2-15 Computed and measured relations between top displacement and base shear
(Moehle and Diebold, 1984)......................................................................................137
Fig. 2-16 Test structure (Hayes, Foutch, and Wood, 1999) .......................................................137
Fig. 2-17 Second-story drift comparison for El Centro simulations on bare model
(Hayes, Foutch, and Wood, 1999) .............................................................................137
Fig. 3-1(a) RC specimen ............................................................................................................138

ix
Fig. 3-1(b) PT specimen.............................................................................................................138
Fig. 3-2(a) Test specimens .........................................................................................................139
Fig. 3-2(b) Prototype and shake table specimens (Hatched)......................................................139
Fig. 3-3(a) Elevation view of RC specimen (Frame N, S).........................................................140
Fig. 3-3(b) Elevation view of RC specimen (Frame W, C, E)...................................................140
Fig. 3-4(a) Details of column.....................................................................................................141
Fig. 3-4(b) Column section ........................................................................................................141
Fig. 3-5(a) Details of footing .....................................................................................................142
Fig. 3-5(b) Footings ...................................................................................................................142
Fig. 3-6(a) Top slab reinforcement of RC specimen..................................................................143
Fig. 3-6(b) Bottom slab reinforcement of RC specimen............................................................143
Fig. 3-7(a) Details of top reinforcement (RC specimen, Exterior connection)..........................144
Fig. 3-7(b) Details of top reinforcement (RC specimen, Interior connections) .........................144
Fig. 3-8 Details of bottom reinforcement (RC specimen) .........................................................145
Fig. 3-9(a) Exterior connection..................................................................................................145
Fig. 3-9(b) Interior connection...................................................................................................145
Fig. 3-9(c) Slab reinforcement of RC specimen ........................................................................145
Fig. 3-10(a) Slab - column critical sections ...............................................................................146
Fig. 3-10(b) Stud-rails................................................................................................................146
Fig. 3-11 Stress – strain relations of first batch concrete cylinder.............................................146
Fig. 3-12 Test specimens ...........................................................................................................146
Fig. 3-13(a) Elevation view of PT specimen (Frame N, S) .......................................................147
Fig. 3-13(b) Elevation view of PT specimen (Frame E, C, W) .................................................147
Fig. 3-14(a) Tendon arrangement of PT specimen ....................................................................148
Fig. 3-14(b) Plan views of bonded reinforcement (PT specimen) .............................................148
Fig. 3-15(a) Overview of tendon arrangement (PT specimen) ..................................................149
Fig. 3-15(b) Interior connection (PT).........................................................................................149
Fig. 3-15(c) Edge connection (PT).............................................................................................149
Fig. 3-16(a) Anchor plate and edge tension bar.........................................................................149
Fig. 3-16(b) Anchors..................................................................................................................149

x
Fig. 3-17(a) Details of interior connection (PT specimen) ........................................................150
Fig. 3-17(b) Details of edge connection (PT specimen) ............................................................150
Fig. 3-18(a) Tendon layouts (E-W direction, See Fig. 3-12) .....................................................151
Fig. 3-18(b) Tendon layouts (N-W direction, See Fig. 3-12) ....................................................151
Fig. 3-19 Donut-shaped load cells .............................................................................................152
Fig. 4-1 Moving PT specimen with bracing system ..................................................................153
Fig. 4-2 Embedded rods .............................................................................................................153
Fig. 4-3 Lifting Anchor ..............................................................................................................153
Fig. 4-4 Baseplates for attachment of specimens.......................................................................153
Fig. 4-5 Plans of baseplates........................................................................................................154
Fig. 4-6 Footing attachment .......................................................................................................154
Fig. 4-7 Footing anchorage ........................................................................................................154
Fig. 4-8 Installation of lead-weights ..........................................................................................154
Fig. 4-9 Layouts of lead-weights................................................................................................155
Fig. 4-10 A pair of steel rollers ..................................................................................................155
Fig. 4-11 Steel pads and rubber pads .........................................................................................155
Fig. 4-12 Wooden dowels ..........................................................................................................156
Fig. 4-13 Re-shoring of PT slabs ...............................................................................................156
Fig. 4-14(a) Strain gauges ..........................................................................................................156
Fig. 4-14(b) Strain gauges applied with M-coat J-3 ..................................................................156
Fig. 4-15 Locations of strain gauges attached on slab reinforcement ........................................156
Fig. 4-16 Strain gauges on top bars (RC specimen, FL1, FL2) .................................................157
Fig. 4-17 Strain gauges on bottom bars (RC specimen, FL1, FL2) ...........................................157
Fig. 4-18 Strain gauges on column bars and stud-rails (RC specimen, Frame N) .....................158
Fig. 4-19 Strain gauges on slab bars and slab surface (PT specimen, FL1, FL2) ......................158
Fig. 4-20 Strain gauges on bars, stud-rails, and column surface
(PT specimen, Frame N) ............................................................................................159
Fig. 4-21 Tri-axial load cell .......................................................................................................159
Fig. 4-22 Locations of donut shaped load cells..........................................................................160
Fig. 4-23 Donut-shaped load cell for tendons............................................................................160

xi
Fig. 4-24 Accelerometers ...........................................................................................................160
Fig. 4-25 Instrumentation on slab bottom and table (RC specimen, FL1).................................161
Fig. 4-26 Instrumentation on slab top (RC specimen, FL1) ......................................................161
Fig. 4-27 Instrumentation on slab top and bottom (RC specimen, FL2) ...................................162
Fig. 4-28 Instrumentation on slab top and bottom (PT specimen, FL1) ....................................162
Fig. 4-29 Instrumentation on slab top and bottom (PT specimen, FL2) ....................................163
Fig. 4-30 Instrumentation (RC specimen, Frame N)..................................................................163
Fig. 4-31 Instrumentation (RC specimen, Frame S) ..................................................................164
Fig. 4-32 Instrumentation (PT specimen, Frame N) ..................................................................164
Fig. 4-33 Instrumentation (PT specimen, Frame S)...................................................................165
Fig. 4-34 DCDTs mounted on top and bottom of slab ..............................................................165
Fig. 4-35 DCDTs mounted on each side of column ..................................................................165
Fig. 4-36 DC displacement transducer.......................................................................................166
Fig. 4-37 Spring potentiometer ..................................................................................................166
Fig. 4-38 SPs installed between footing and steel plate.............................................................167
Fig. 4-39(a) Rotation of footing plane monitored using SPs .....................................................167
Fig. 4-39(b) Locations of SPs monitoring footing rotation (PT specimen) ...............................167
Fig. 4-40 Table acceleration histories of each test (RC specimen)............................................168
Fig. 4-41 Table acceleration histories of each test (PT specimen) ............................................169
Fig. 4-42 Response spectra of each test .....................................................................................170
Fig. 4-43 Time-compressed ground motion...............................................................................171
Fig. 5-1 Process used to compute base shear and top relative displacement .............................172
Fig. 5-2 Footing rotation vs moment relations...........................................................................172
Fig. 5-3(a) Base shear vs top relative displacement relations (RC specimen)...........................173
Fig. 5-3(b) Base shear vs top relative displacement relations (PT specimen) ...........................173
Fig. 5-4 Moment vs curvature relations at the column base ......................................................174
Fig. 5-5 Damage at the column base (RC specimen).................................................................174
Fig. 5-6 Process used to compute moment from measured curvature .......................................175
Fig. 5-7 Column moments from independent measures ............................................................176
Fig. 5-8 Column curvatures from independent measures ..........................................................177

xii
Fig. 5-9 Strain gauges and displacement gauges used to measure slab curvature .....................177
Fig. 5-10(a) Slab strain gauge data located symmetrically (RC, FL1NW, Top bars) ................178
Fig. 5-10(b) Slab strain gauge data located symmetrically (RC, FL1NW, Bottom bars) ..........178
Fig. 5-11 Slab curvature variations of exterior connections (RC, FL1NC-w)...........................178
Fig. 5-12 Slab strain gauge data (RC specimen)........................................................................179
Fig. 5-13(a) Slab strain gauge data (RC, FL1NC, Top bars) .....................................................179
Fig. 5-13(b) Slab strain gauge data (RC, FL1NC, Bottom Bars)...............................................179
Fig. 5-14(a) Damage on top slab at the end of testing (RC, FL1SE) .........................................180
Fig. 5-14(b) Damage on bottom slab at the end of testing (RC, FL2SW) .................................180
Fig. 5-15 Flexural cracks across the full width at the end of testing .........................................180
Fig. 5-16(a) Damage on top slab at the end of testing (RC specimen, FL1SC).........................181
Fig. 5-16(b) Damage on bottom slab at the end of testing (RC specimen, FL1NC)..................181
Fig. 5-17 Damage on top and bottom slabs at the end of testing (PT specimen).......................181
Fig. 5-18 Torsional cracks on top slab at the end of testing (PT specimen, FL2SE).................182
Fig. 5-19 Substantial concrete spalling on west edge at the end of testing (RC, FL1SW) ........182
Fig. 5-20(a) Curvature diagrams (RC specimen, FL2NW)........................................................183
Fig. 5-20(b) Curvature diagrams (RC specimen, FL2NE).........................................................183
Fig. 5-21(a) Curvature diagrams (PT specimen, FL2NC-w) .....................................................184
Fig. 5-21(b) Curvature diagrams (PT specimen, FL2NC-e) ......................................................184
Fig. 5-22(a) Curvature diagrams (PT specimen, FL1NC-w) .....................................................185
Fig. 5-22(b) Curvature diagrams (PT specimen, FL2NC-e) ......................................................185
Fig. 5-23 Slab moment diagrams within c + 3h at the first yield of slab reinforcement –
RC Specimen .............................................................................................................186
Fig. 5-24 Slab moment diagrams within column strip at the first yield of slab reinforcement –
PT Specimen ..............................................................................................................186
Fig. 5-25 Process used to compute slab rotation (RC specimen)...............................................187
Fig. 5-26 Process used to compute slab rotation (PT specimen) ...............................................187
Fig. 5-27 Drift ratio capacity with the existing database (See Table 5-2 and 5-3) ....................188
Fig. 5-28 Post-tensioning force distribution (at the completion of first post-tensioning)..........188
Fig. 5-29 Post-tensioning force distribution (at the completion of third post-tensioning) ........189

xiii
Fig. 6-1 Analytical model for displacement...............................................................................190
Fig. 6-2 Inelastic model for rigid plastic springs .......................................................................190
Fig. 6-3 Crack coefficient for columns ......................................................................................191
Fig. 6-4 Top relative displacement vs base shear comparisons for elastic model .....................192
Fig. 6-5 Top relative displacement comparisons for elastic model ...........................................193
Fig. 6-6 Top relative displacement comparisons at peak (RC)..................................................194
Fig. 6-7 Top relative displacement comparisons at peak (PT)...................................................195
Fig. 6-8 Top relative displacement vs base shear comparisons for fiber element model ..........196
Fig. 6-9 Top relative displacement comparisons at peak for fiber element model ....................197
Fig. 6-10 Push-over curves for top relative displacement vs base shear ...................................198

xiv
1. Introduction
1-1. Introduction

Common design practice in the western United States allows the structural engineer to

design a building with a lateral-force-resisting-system (LFRS) and a non-participating

system or gravity-force-resisting system (GFRS). The elements designated to be part of

the LFRS are proportioned to resist the entire design seismic forces and provide

sufficient stiffness to limit the lateral drift to acceptable levels. The elements designated

to be part of the GFRS are proportioned assuming that they do not contribute to the

seismic resistance, that is, these elements are designed to resist gravity forces only. The

ability of the GFRS to support the gravity loads when subjected to the design lateral

deformations must be checked, commonly referred to as a deformation compatibility

check.

For buildings less than approximately ten stories, it is common practice on the west

coast of the U.S. to use a perimeter special moment frame with an interior slab - column

gravity frame. For taller buildings (10 to 25 stories), use of a core wall system to

provide lateral strength and stiffness, and a slab-column gravity frame is fairly common.

This system has also been used on buildings with as many as 40 stories (Klemencic,

ACI Fall Convention, Phoenix, AZ, Oct., 2002). The use of a post-tensioned floor slab

also is common, as it allows for longer spans for the gravity frame and is easy to

construct. The post-tensioning strands are typically banded in one direction, and

distributed in the other direction.

1
Given the longer spans for the post-tensioned floor system, drop panels may be

necessary due to the higher shear stresses that develop at the slab - column interface,

although the use of shear reinforcement within the slab adjacent to the column has

emerged as the preferred solution to addressing the high slab - column connection shear

stresses. Shear reinforcement may take the form of stirrups (e.g., Robertson et al.,

2002), so-called shear bands (Pilakoutas and Ioannou, 2000), or so-called stud-rails

(e.g., Elgabry and Ghali, 1987; Fig. 1-1). The use of stud-rails is very common because

they are easy to place and test results have been published that have shown them to be

effective. The use of shear reinforcement increases the shear strength of the slab -

column connection (Vn = Vc + Vs), where Vc and Vs are the nominal shear strength

provided by the concrete and shear reinforcement, respectively. The increased shear

strength typically eliminates the need for a drop panel or column capitol, resulting in

reduced construction costs. Shear reinforcement has also been shown to increase the

deformation capacity or ductility of slab - column connections.

Design of slab-column connections is covered in the ACI 318 Building Code

(“Building”, 2002; see Section 21.12.6). For combined lateral and gravity loads,

sufficient flexural reinforcement must be placed within the column strip to resist the

unbalanced moment, and, for an interior connection, at least one-half of this

reinforcement must be placed within c2 + 3h, where c2 is the column dimension

perpendicular to the direction of the applied loads and h is the slab thickness. In

addition, the ability of the slab - column connection to transfer the unbalanced moment

to the column must be checked. Transfer of the unbalanced moment is assumed to occur

through two mechanisms, flexure and eccentric shear. Reinforcement to resist the

fraction of the unbalanced moment transferred in flexure γfMunb (e.g., 60%) must be

2
placed within c2 + 3h. The remaining fraction of the unbalanced moment (1-γf )Munb =

γvMunb (e.g., 40%) is transferred in eccentric shear on the slab critical section, defined to

extend d/2 from the column face (Fig. 1-2). The resulting maximum shear stress on the

critical section is calculated as (Fig. 1-2):

Vg γ vM uc
vu = v direct + vunb = + (1-1)
bo d Jc

Where vdirect is the gravity shear stress on the critical section, bo = 2[(c1+d) + (c2+d)], d

is the effective section depth, Vg is the gravity force to be transferred from the slab to

the column, vunb is the shear stress on the critical section due to the unbalanced moment

being transferred in eccentric shear, c is the distance from the centroid of the critical

section to the perimeter of the critical section, and Jc is the polar moment of inertia of

the critical section (see Park and Gamble, 2000; Section 10.3.3). The combined shear

stress vu on the critical section is found by adding the direct shear stress and the

eccentric shear stress, which must be less than the nominal shear stress capacity of the

critical section, reduced by φ, that is:

vu ≤ φv n = φ (vc + v s ) (1-2)

Where vc and vs are the shear resistance provided by concrete and shear reinforcement,

respectively.

Experimental studies (e.g., Pan and Moehle, 1989, 1992; Moehle 1996) have shown that

the magnitude of the gravity shear stress on the critical section significantly influences

the drift level at which a connection punching failure occurs (Fig. 1-3). For gravity

shear stress ratios greater than 0.4, tests of slab-column connections reveal little

displacement ductility capacity; therefore, ACI 318 (Building, 2002) places a limit of

3
0.4φVc on the concrete shear strength of the critical section (S21.12.6.8). The three data

points plotted in Fig. 1-3 for isolated, post-tensioned, slab - column connections

subjected to slowly varying drift cycles implies increased drift capacity prior to

observed punching failures. However, due to the longer spans, post-tensioned slab -

column floor systems tend to be more flexible then systems without post-tensioning and

also tend to have higher gravity shear stress ratios at slab - column connections.

A representative punching failure for reinforced concrete construction without shear

reinforcement is shown in Fig. 1-4. The addition of “stud-rail” shear reinforcement is

superimposed to show how the addition of shear reinforcement would cross the failure

plane and act to increase slab - column shear capacity. Elgabry and Ghali (1987)

developed design recommendations for interior slab - column connections reinforced

with stud-rails based on tests of five isolated reinforced concrete slab - column

connections tested primarily under monotonic loads. Additional tests have been

conducted to address design provisions for edge and corner connections with stud-rails

(Mortin and Ghali, 1991; Hammill and Ghali, 1994) and the seismic design of interior

connections subjected to cyclic loading (Megally and Ghali, 2000).

1-2. Summary

US west coast design practice has evolved such that the lateral strength and stiffness are

provided with either a perimeter frame or a core wall, or both (dual system). A slab -

column frame is commonly used for gravity loads. Post-tensioned floor slabs are also

common to increase span lengths, and the use of shear reinforcement in the form of

stud-rails is used to increase the shear strength of the slab - column connection to allow

for a thinner slab or to eliminate the need for drop panels or column capitols.

4
Limited test data exist from which to assess the appropriateness of current design

practice. The performance of stud-rail shear reinforcement is based on tests of isolated

connections subjected to monotonic and cyclic loading (See Robertson et al., 2002; pp.

612). Data for post-tensioned connections are limited to three isolated slab - column

connections tested at UC Berkeley (Qaisrani, 1993). As well, test results conducted on

isolated specimens can be influenced by boundary conditions, and tests conducted under

static loading conditions may overstate damage due to excessive crack propagation due

to the application of sustained loads relative to a dynamic test. Given the prominent use

of post-tensioned slab - column systems with stud-rail shear reinforcement, an

experimental study for dynamic loading was undertaken to assess system performance.

The experimental program undertaken consisted of the design, fabrication, and shake

table testing of two, approximately one-third scale, sub-assemblies on the shake table at

the UC Berkeley EERC Richmond Field station. Both specimens consisted of 2 × 2 bay,

two story, slab - column frames (Fig. 1-5). One specimen consisted of a reinforced

concrete slab - column frame whereas the other specimen consisted of a post-tensioned

slab - column frame. Shear reinforcement (stud-rails) were used to increase the nominal

shear strength of slab column connections for both specimens. The specimens were

designed such that yielding of slab flexural reinforcement was expected prior to shear

failure within the shear reinforced region of the slab - column connections.

1-3. Report Organization

Following this introduction, a literature review is presented in Chapter 2, followed by a

detailed description of the test specimens and construction in Chapter 3. Test setup and

instrumentation of the specimens is detailed in Chapter 4. The shake table motions used

5
in the test program are also documented in Chapter 4. Preliminary results obtained from

the testing program are presented in Chapter 5, followed by preliminary results obtained

from analytical modelling studies (Chapter 6). Preliminary conclusions are presented in

Chapter 7.

6
2. Literature Review

Available information on the lateral-load stiffness, strength, and ductility of slab -

column connections are reviewed in the following subsections. This background

information serves as the foundation for future chapters.

2-1. Lateral-load Stiffness

The lateral-load stiffness of a slab - column frame is often represented using an

“equivalent” slab-beam to account for the flexural stiffness of the slab. One approach,

referred to as an effective beam width model, provides a simple and reasonably accurate

means to model the lateral-load stiffness of slab in a slab - column frame (Fig. 2-1a). An

alternative approach, the equivalent frame model, represents the stiffness of the slab -

column frame using an equivalent beam in series with a spring representing the

torsional stiffness of the slab adjacent to the connection region (Fig. 2-1b).

As shown in Fig. 2-1, the three-dimensional system can be modeled as two-dimensional

frame using an effective slab width αl1 and a conventional column, where the α-factor

is derived using elastic plate theory to result in an equivalent slab width with uniform

rotation across the effective slab width that yields the same rotation stiffness as the

original system with non-uniform rotation (Pecknold, 1975). For a square slab, typical
c1
values of the effective slab width αl1 range from (0.3 to 0.6)l1, for ratios of between
l1
0.05 to 0.10, where c1 is the dimension of rectangular column parallel to loading

direction and l1 is the length of span parallel to loading direction (Allen and Darvall,

1975).

7
Based on finite element analysis, Banchik (1987) introduced the following effective

beam width factors to define the uncracked slab moment-rotation stiffness:

 l  1
b =  5c1 + 1  : for interior frame lines (2-1)
 4  1 −ν 2

 l  1
b =  3c1 + 1  : for exterior frame lines (2-2)
 8  1 −ν 2

Where l2 is the length of span transverse to l1, and µ is Poisson’s ratio. Equations (2-1)
c2
and (2-2) apply only for the column aspect ratio between 1/2 and 2, and slab aspect
c1
l2
ratio between 2/3 and 3/2, where c2 is the length of column transverse to c1.
l1

Hwang (1989) evaluated the effective beam width model suggested by Banchik (1987),

by reviewing existing analytical and experimental research, as well as conducting a test

on a 4/10 scale, 3 × 3 bay structure with both square and rectangular columns. Hwang

(1989) reported that (2-1) and (2-2) represented the experimental results reasonably well.

Tests conducted on four 7/16 scale interior connections by Qaisrani (1993) and two 3/7

scale edge connections by Martinez (1993), revealed that the relations suggested by

Banchik (1987) were appropriate for both interior and edge connections of post-

tensioned slabs.

Modifications to the relations used to estimate the effective slab width (e.g., (2-1) and

(2-2)), are required to account for the influence of cracking. Vanderbilt and Corley

(1983) recommended a stiffness reduction multiplier of 1/3 for the equivalent frame

model. Subsequently, Moehle and Diebold (1984) recommended a value between 1/3

and 1/2 for both the equivalent frame model and the effective beam width model. To

address cracking effects quantitatively, the following general expression was developed

8
by Hwang (1989) using elastic plate theory and then verified using experimental data to

account for the influence of cracking.

M 
0.75
 M 
0.75
I
β =  cr  + 1 −  cr   cr (2-3)
 Ma    M a   I g

Where Mcr is the product of slab unit cracking moment and the column perimeter (πd),

Ma is the applied moment, Icr is the moment of inertia of cracked concrete section, and

Ig is the gross moment of inertia of gross concrete section.

For slab - column connections designed according to the ACI 318-83 (“Building”,

1983) with a factor of safety of 1.4 on transfer moment, the approximate expressions of

stiffness reduction due to cracking, β, can be written as (Hwang, 1989):

c  L  1
β = 5 − 0.1 − 1 ≥ , where L = service live load in units of kPa (2-4)
l  1.915  3

Equation (2-4) can be approximated as (Hwang, 1989):

c 1
β =4 ≥ (2-5)
l 3

Where c is the dimension of a square column, l is the length of a square slab panel. For

post-tensioned interior and edge connections, Martinez (1993) and Qaisrani (1993)

suggested values of β equal to 1/3 and 1/2, respectively. Slab reinforcement for the tests

conducted by Martinez (1993) and Qaisrani (1993) consisted of banded tendons and

bonded top reinforcement placed within c2 + 3h to control cracking and reduce stiffness

degradation for the post-tensioned slab-column connections, where h is the slab

thickness (Smith and Burns, 1974; Burns et al., 1977; 1985; Kosut et al., 1985; Martinez,

9
1993).

The equivalent frame model was first proposed by Peabody (1948) for the evaluation of

slab - column frames subjected to gravity loads. In the equivalent frame model, slab-

beams with length equal to the full span l2 are connected to columns by a pair of

transverse-torsion members as shown in Fig. 2-1(b). The rotational connection stiffness

is determined as the combined stiffness of the slab-beam with dimensions c2 × h, and

the stiffness of the transverse torsion member (Kt). The torsional stiffness (Kt) in (2-6) is

determined from the torque – rotation relation (ACI 318-02, “Building”, 2002; See

Section 13.7.5).

9 E cs C
Kt = ∑ 3
(2-6)
 c 
l 2 1 − 2 
 l2 

 x  x3 y
Where Ecs is the modulus of elasticity of the slab concrete, C = ∑
 1 − 0 . 63 
y  3
is

the cross-sectional constant to define torsional properties, l2 is the length of span

perpendicular to the direction that moments are being determined, c2 is the dimension of

rectangular column perpendicular to the direction that moments are being determined.

The equivalent frame model has been empirically developed and calibrated with a

substantial number of tests.

Vanderbilt and Corley (1983) extended the equivalent frame model for application to

combined gravity and lateral loading, and Trongtham and Hawkins (1977) revealed that

the equivalent frame model is valid for post-tensioned slabs subjected to gravity and

lateral loads provided the tendons provide sufficient compression to eliminate cracking.

10
Luo et al. (1994; 1995) proposed an effective beam width model combined with the

equivalent frame model, based on column and slab aspect ratios and the magnitude of

the gravity load (Fig. 2-2). Figure 2-2 schematically depicts an effective beam width

factor for an interior connection, αi, as based on recommendations by Pecknold (1975),

and an effective beam width factor for an edge connection, αe, by considering the

torsional stiffness of slabs adjacent to the edge connections for an equivalent frame

model. The effective beam width factors suggested by Luo et al. (1994; 1995) are

written as:

c2
1.02( )
l2
αi = (2-7)
l c c c
0.05 + 0.002( 1 ) 4 − 2( 1 ) 3 − 2.8( 1 ) 2 + 1.1( 1 )
l2 l1 l1 l1

Kt
αe = (2-8)
Kt + Ks

(4 E cs I )
Where Ks = is the flexural stiffness of slab. Equation (2-7) applies only for 0.5
l1
c2 l
≤ ≤ 2.0 and 0.5 ≤ 2 ≤ 2.0.
c1 l1

Based on the review of 40 interior connection tests, Luo and Durrani (1995) suggested

the following reduction factor χ be introduced for both interior and edge connections.

Vg
χ = 1 − 0.4 (2-9)

(1 / 3) Ac fc

Where Vg is the gravity force to be transferred from the slab to the column in unit of

MPa, Ac is the area of slab critical section specified in ACI 318-02 (“Building”, 2002),

and fc’ is the compressive strength of concrete. Use of the reduction factor χ results in

11
an effective moment of inertia of the equivalent slab width as:

M 
3
 M 
3

I e =  cr  I g + 1 −  cr   I cr (2-10)
 Ma    M a  

1
Where Ig = χα i l 2 h 3 and represents the gross moment of inertia of effective slab
12
width, Icr is the moment of inertia of cracked effective slab width, Mcr is the cracking

moment of effective slab width, and Ma is the applied moment.

2-2. Lateral-load Strength

Several failure mechanisms can develop in a slab - column frame. One failure mode

involves the development of a flexural yield line(s) across the full width of the slab. In

this case, the slab - column connection is strong enough to transfer the unbalanced

moment and shear that develop due to the development of the flexural yield line(s). If

the slab - column connection is not capable of resisting the unbalanced moment and

shear that develop for a flexural yield line(s) across the slab, then the slab – column

connection fails first. Where slab - column connection failure occurs, two failure modes

are possible: (1) yielding of slab flexural reinforcement followed by punching shear

failure, or (2) punching shear failure prior to flexural yielding of the slab flexural

reinforcement. Models used to assess failure modes and the capacity of the slab -

column connections to transfer unbalanced moment and shear from the slab to the

column are described in the following subsections.

2-2-1. Yield Line Failure

Figure 2-3 depicts flexural yield lines across the full width of a slab - column frame for

12
an interior and exterior connection. For a flexural yield line to develop across the entire

width of a slab, the slab - column connection must have sufficient capacity to transfer

the unbalanced moment and shear at the slab - column connection when the yield line

forms. Therefore, models to assess slab - column connection capacity are needed.

2-2-2. Eccentric Shear Stress Model

The eccentric shear stress model, introduced by Di Stasio and Van Buren (1960), is

commonly used to assess the capacity of a slab - column connection to transfer

unbalanced moment and shear from the slab to the column. Hanson and Hanson (1968)

conducted experimental studies to verify the eccentric shear stress model, and a number

of subsequent studies have been conducted to assess the validity of the model and to

extend or improve the model (Hawkins, 1971; Hawkins and Corley, 1971; Islam and

Park, 1976; Moehle and Diebold, 1984; Zee and Moehle, 1984; Robertson and Durrani,

1990; Pan and Moehle, 1992).

In the eccentric shear stress model, unbalanced moment and shear in the slab are

transferred to the column by: (1) direct shear and eccentric shear acting on a critical

section, and (2) flexural yielding over a slab transfer width. The unbalanced moment

and shear are determined from analysis using equilibrium requirements at the

connection. The direct shear and unbalanced moment that must be transferred are

determined from vertical and moment equilibrium requirements, respectively, for all the

members framing into a connection.

According to ACI 318-02 (“Building”, 2002), a fraction of unbalanced moment, γfMunb,

is transferred by flexure over a transfer width of c2 + 3h centered on the column. The

13
fraction γf is determined using (13-1) of ACI 318-02 (“Building”, 2002) as:

1
γ f = (2-11)
2 b1
1+
3 b2

Where b1 is the width of the critical section parallel to loading direction, and b2 is the

width of the critical section transverse to b1. Sufficient reinforcement must be placed

within c2 + 3h to resist γfMunb. This flexural reinforcement may consist of top or bottom

reinforcement, or some combination of top and bottom reinforcement, depending on the

connection configuration and the direction of the unbalanced moment. The flexural

reinforcement within the transfer width must be anchored to develop the yield stress of

the reinforcement at the critical section of the transfer width (Fig. 2-4). Some

researchers have recommended increasing the flexural transfer width. Hawkins et al.

(1989) suggested a flexural transfer width of c2 + 4h based on a review of results from

tests conducted on 36 interior slab-column connections. Hwang (1989) suggested that

the flexural transfer width could be increased to c2 + 5h.

Once the fraction of the unbalanced moment has been assigned to the flexural transfer

width, the remaining fraction of the unbalanced moment (γv = 1 - γf) must be transferred

through eccentric shear. Details associated with shear transfer on the critical section are

depicted in Fig. 2-5. The stresses due to direct shear and eccentric shear on the critical

section and the shear strength of the connection for non-prestressed connections are

computed as:

Vu γ v M unb cc φVn
vu = ± ≤ (2-12)
Ac Jc bo d

14
 ′ ′ 
 2( 4)  f c bo d  α s d  f c bo d 1 ′
φVn = φVc = min  2 +  ,  + 2  , f c bo d  (MPa)  (2-13)
 β c  6  bo  12 3 
 

Where Vu is the factored shear force to be transferred from the slab to the column, Munb

is the factored unbalanced moment to be transferred from the slab to the column, cc is

the distance from the centroid of the critical section to the perimeter of the critical

section, Jc is the polar moment of inertia of the critical section (see Park and Gamble,

2000; Section 10.3.3), bo = 2[(c1+d) + (c2+d)] is the perimeter of the critical section, d is

the effective section depth, βc is the ratio of long side to short side of the column, αs is

40 for interior columns, 30 for edge columns, 20 for the corner columns, and φVn is the

nominal shear capacity of the connection reduced by the capacity reduction factor in

units of MPa. For two-way prestressed slabs, φVn is computed as:

 ′ Vp 
φVn = φVc = φ  β p f c + 0.3 f pc +  (MPa) (2-14)
 bo d 

Where βp is the smaller of 0.29 or (αsd/bo + 1.5)/12, fpc is the average compressive

stress in concrete due to the effective prestress force only, Vp is the vertical component

of effective prestress force at section.

For slab - column connections with shear reinforcement, the shear strength of the

connection must be checked at two locations: (1) d/2 from the column face within the

shear-reinforced region, and (2) d/2 outside the shear-reinforced region (Fig. 2-6).

Calculations required for the critical section outside the shear-reinforced region are

essentially the same as noted previously, except for the change in the geometry of the

critical section, and the shear strength φVn at a critical section outside the shear-

15
reinforced zone is computed as:

1 ′ 
φVn = φ  f c bo d  (MPa) (2-15)
6 

Where φVn is the nominal shear capacity of the connection reduced by the capacity

reduction factor.

Within the shear-reinforced region, the nominal shear strength of the slab critical section

is enhanced as (ACI 318-02, “Building”, 2002):

Vu γ v M unb cc φ (Vc + Vs )
vu = ± ≤ (MPa) (2-16)
Ac Jc bo d

Where Vc is the nominal shear capacity provided by concrete, which cannot exceed

1 ′ Av f y d
f c bo d , Vs = is the nominal shear strength provided by shear reinforcement,
6 s
Av is the area of shear reinforcement within a distance s, fy is the yield strength of shear

reinforcement, s is the spacing of shear reinforcement. The combined shear strength, Vc


1 ′
+ Vs cannot exceed f c bo d .
2

Studies by Smith and Burns (1974), Burns and Hemakom (1977; 1985), Trongtham and

Hawkins (1977), Kosut et al. (1985), Foutch et al. (1990), Long and Cleland (1993),

Martinez (1993), and Qaisrani (1993) indicate that the eccentric shear stress model

gives reasonable results for post-tensioned slabs (Table 2-1), as ratios for most cases are

close to unity. In Table 2-1, the shear strength was computed using (2-14); however, the

contribution of Vp was neglected as suggested in the commentary of ACI 318-02

(“Building”, 2002) to yield a conservative estimate of concrete shear strength. The tests

of isolated, post-tensioned connections subjected to cyclic or monotonic lateral loading

16
were conducted by Trongtham and Hawkins (1977), Foutch et al. (1993), Martinez

(1993), and Qaisrani (1993), whereas Smith and Burns (1974), Burns et al. (1977; 1985),

Kosut et al. (1985), and Long and Cleland (1993) tested various subassemblies of post-

tensioned flat plate systems under uniform gravity load only. More details are provided

in Chapter 2, Section 5.

ACI 318-02 (“Building”, 2002) allows the fraction of the unbalanced moment being

transferred in flexure to be increased in some cases. The value γf may to be increased up

to 1.0 for edge connections and (1.25)γf for interior connections, if the factored gravity

shear ratio Vu/φVc is less than 0.75 and 0.4, respectively, provided that the reinforcing

ratio ρ within c2 + 3h does not exceed 0.375ρb.

Alternative recommendations for defining parameters associated with the shear and

unbalanced moment transfer are available. ACI 352.1R-89 recommends that the

strength of reinforced concrete slab - column edge connections be defined as the lesser

of either three-quarters of the direct shear strength without the transfer of unbalanced

moment or the flexural strength within a transfer width of c2 + 2c1. Based on a review of

27 connection tests, Moehle (1988) reported that there is no evidence of interaction

between shear and applied normal moment for an edge connection at failure provided

the ratio of the applied gravity shear to the shear strength of the critical section is less

than 0.75 (Fig. 2-7). Tests of two one-half scale subassemblies conducted by Robertson

and Durrani (1990) also revealed little interaction between shear and unbalanced

moment transfer for exterior connections. Robertson and Durrani (1990) recommended

that the fraction of unbalanced moment transferred by flexure for interior connections

could be varied between 0.6 and 0.8 without impacting overall connection behavior.

17
2-3. Lateral-load Ductility

Slab - column frames are not assumed to provide lateral-force resistance on the west

coast of the United States. However, due to deformation compatibility requirements

between the lateral-force resisting system and the slab - column frame, the slab -

column frame must be capable of sustaining the gravity loads under the expected lateral

deformations imposed on the system from the design earthquake. In areas of low-to-

moderate seismic risk, slab - column frames, referred to as intermediate moment frames

in ACI 318-02 (“Building”, 2002), may be used as a lateral-force resisting system.

Therefore, the lateral deformation capacity of a slab - column frame is an important

consideration.

The influence of the direct gravity shear stress on the lateral-load ductility of a slab -

column connection was recognized by Kanoh and Yoshizaki (1979). To ensure some

level of ductility in slab - column frames, Hawkins and Mitchell (1979) recommended


limiting the connection shear stress 0.17 f c MPa , if the flexural reinforcement was

expected to yield. ACI 352.1R-89 recommends the gravity shear ratio Vg/Vc not exceed

0.4 to avoid punching failures which tend to occur for lateral drift ratios exceeding 1.5%.

The ACI 318-02 code (“Building”, 2002) limits the factored gravity shear to 0.4φVc in

21.12.6.8.

Data from 26 tests are presented in Fig. 2-8 to compare the observed drift ratio at

punching failure versus the gravity shear ratio Vg/Vc. A clear trend, with decreasing

drift ratio at punching failure as the gravity shear ratio is increased, is observed. Given

that reinforced concrete slab - column frames tend to yield at approximately 1.5%

18
lateral drift (Pan and Moehle, 1989), a gravity shear ratio less than 0.4 is generally

required to ensure displacement ductility ratios greater than one. Based on these data,

Aschheim and Moehle (1995) suggested that the relationship between the connection

rotational capacity be expressed as a linear function of gravity shear ratio (Fig. 1-3).

2-4. Post-punching Behavior

Although loss of moment strength and rotational stiffness due to punching failure is a

concern, the primary issue associated with post-punching behavior of slab - column

connections is to avoid progressive collapse, or “pancaking” (Bullock’s, Northridge

Earthquake Reconnaissance Report, 1996) by providing a mechanism for supporting the

gravity loads after punching failure. This is typically accomplished by the use of

continuous bottom reinforcement through the connection region. According to ACI

352.1R-89, the required area of continuous bottom reinforcement Asm is equal to


0.5ω u l1l 2
, based on an assumed angle of 30 degrees from horizontal for the bottom
φf y

reinforcement after punching, and where ωu is the uniformly distributed design load, l1

and l2 are the center to center span in orthogonal directions, φ = 0.9, and fy is the yield

stress of steel. It is recommended that only bottom reinforcement be considered in this

calculation, although there is some evidence that top reinforcement is also effective

(Pan and Moehle, 1989).

According to ACI 423.3R-96, at least two tendons should be placed through the column

cage in each direction. These are effective in supported a slab after punching failure,

nevertheless, some continuous bottom steel at interior connections of post-tensioned

slab - column frames is desirable (Moehle, 1996). Therefore, structural integrity

19
reinforcement (continuous bottom reinforcement) as specified in ACI 352.1R-89 also

should be used for post-tensioned interior connections.

2-5. Post-Tensioned Flat Plate

In the prior sections, general details concerning the behavior of slab - column frames

were discussed with some information on post-tensioned slabs included. Additional

details for post-tensioned slab - column frames are reviewed in the paragraphs that

follow.

Smith and Burns (1974), Burns and Hemakom (1977; 1985), and Kosut et al. (1985)

conducted tests on post-tensioned slabs to assess behavior for gravity loads. Three

different specimen configurations were tested: (1) three isolated specimens reinforced

with various amounts of bonded mild reinforcement, (2) two 3 × 3 bay specimens, one

specimen with 70% of the tendons within the column strip and the other specimen with

a banded arrangement of unbonded tendons, and (3) one 2 × 2 bay specimen with a

banded arrangement of unbonded tendons.

Test results indicated that bonded reinforcement was effective in increasing the load

carrying capacity and ductility, as well as reducing maximum crack widths. Bonded top

reinforcement was found to provide effective crack control if placed within a distance of

1.5 times the slab thickness from the face of the columns. Use of banded tendons over

the column improved connection behavior by providing high local compressive stresses

in the connection region, although the tendon stresses at the ultimate load capacity of

the specimens were less than those predicted by ACI 318-77, Eq. 11-13 (“Building”,

1977).

20
Trongtham and Hawkins (1977) subjected six, full-scale, post-tensioned slab - column

connections with both top and bottom bonded reinforcement to cyclic loading. The ends

of the columns extending through the slab were pinned, and the slab edges were

displaced vertically to apply unbalanced moment to the connections (Fig. 2-9). Of the

six specimens, four were interior connections, one was an edge connection, and one

represented an interior lift slab connection. All interior connections achieved large

deflections, between 3 in. (7.6 cm) and 6 in. (15.2 cm), prior to punching failure,

followed by rapid loss of capacity and pinching of the load-deformation response.

Yielding of bonded top reinforcement was observed; however, yielding of the post-

tensioning tendons was not observed, providing for some elastic recovery. Yielding of

unbonded post-tensioning reinforcement is unlikely, as the elongation is distributed over

the entire strand. Therefore, given that only bonded bars yield, the ductility of post-

tensioned slab - column connections is controlled by the ratio of strength provided by

the bonded and unbonded reinforcement. Providing balanced quantities of tendons and

bonded reinforcement in the connection region improved hysteretic damping (ACI

423.3R-96).

Lateral load versus deflection relationships for reinforced concrete interior connections

with and without shear reinforcement, and a post-tensioned interior connection, are

plotted in Fig. 2-10 (Hawkins et al.; 1975; 1975; 1977). The three specimens have

similar geometry, bonded reinforcement, and gravity load, although the column sizes

vary slightly. The shear-reinforced slab achieved approximately 1.5 times the lateral

load of the non-shear reinforced slab, and approximately twice the displacement, prior

to punching failure. The post-tensioned specimen was subjected to a greater number of

loading cycles and displayed approximately twice the displacement ductility capacity of

21
the shear-reinforced slab.

Foutch et al. (1990) tested four, two-thirds scale, isolated post-tensioned edge

connections under monotonic loading. Banded tendons were used perpendicular to the

exterior edge of the slab with the column in two of the specimens (S1 and S2), and

parallel to the exterior edge with the column for the other two specimens (S3 and S4).

The banded tendon arrangement used is common practice, as it eases placing and

jacking of the tendons. Eleven tendons in the loading direction were used for specimens

S1 and S2, whereas four tendons were placed in the loading direction of specimens S3

and S4. Moment-deflection relationships for the four specimens are presented in Fig. 2-

11. Results indicate that the stiffness and strength of the specimens prior to failure were

not significant influenced by the tendon arrangement. Substantial ductility was observed

for the specimens S1 and S3, which were subjected to larger ratios of moment and shear

compared to S2 and S4, where punching failure was observed just after yielding. The

test results indicate that shear to moment ratio is an important parameter and influences

the failure type and ductility of the post-tensioned edge connections.

Dilger and Shatila (1989) tested six full-scale, isolated post-tensioned slab - column

edge specimens. Test specimen properties and failure modes are summarized in Table 2-

2. Four of the specimens included shear reinforcement, whereas two did not. All

specimens except one failed in ductile manner. The increases in tendon stress observed

at punching failure in specimens without shear reinforcement was similar to that

observed by Trongtham and Hawkins (1977) and Burns and Hemakom (1985), but

lower than the values reported by Foutch et al. (1990).

Three 7/16 scale interior connections, two 3/7 scale edge connections, and two 3/7 scale

22
corner connections subjected to biaxial lateral loading were tested by Qaisrani (1993)

and Martinez (1993). All connections employed banded tendon layouts in one direction

and the uniformly distributed tendon layouts in the other direction. For one of the two

edge connections, tendons were banded to the slab edge, whereas tendons were banded

parallel to the slab edge for the other two edge connections. All specimens were

subjected to biaxial lateral loading (cloverleaf pattern of lateral drift), and punching

shear failures were observed in all specimens. For the three interior connections, a

sudden drop in lateral load capacity was observed at vector drift ratios of 2.26%, 2.1%,

and 1.8%, for gravity shear ratios of 0.52, 0.66, and 0.72, respectively (Fig. 2-12). The

vector drift ratios at punching failure for the edge connections with gravity shear ratios

of approximately 0.49 and 0.49 were 3.9% and 3.77%, respectively. Both post-tensioned

and reinforced concrete slab - column connections subjected to biaxial lateral loading

displayed greater deterioration in lateral load strength, stiffness, and ductility relative to

specimens subjected to uniaxial lateral loading (Pan and Moehle, 1992; Qaisrani, 1993).

The banded arrangement of tendons did not improve the ultimate strength of the interior

connections tested, which confirms that the ultimate strength of an interior connection is

dependent only on the total flexural reinforcement, and not the distribution and direction

of the reinforcement (Qaisrani, 1993). For edge connections, banded tendons and

reinforcing bars close to slab surface reduce stiffness degradation, increase connection

strength, resulting in greater displacement ductility; however, banded tendons accelerate

strength decay after punching failure due to high local pressure (Martinez, 1993).

2-6. Shake Table Tests of Flat Plate Systems

Moehle and Diebold (1984) conducted a shake table test a of flat plate system. The test

23
structure was one-third scale, with two-stories and three-bays. Spandrel beams were

provided at the perimeter as shown in Fig. 2-13. A gravity shear ratio Vg/Vo of 0.2 was

selected for the interior connections to model approximately the slab self-weight of the

prototype structure. Acceleration histories applied to the shake table were compressed to

account for scale, and were varied from low to high intensity.

The relationship between base shear and top displacement is given in Fig. 2-14 for the

test structure. The relation is relatively linear up to 0.5% top drift ratio and yield is

observed at approximately 1.5% top drift ratio. A lateral drift ratio of 5.3% was

achieved without collapse. Punching at a first floor interior connection occurred at 4.4%

drift. Zee and Moehle (1984) conducted quasi-static tests of individual connections

having the same geometry as the frame tested on the shake table. The isolated

connections behaved similar to the connections in the dynamic test, except that the

isolated interior connections experienced sudden failure at 4% drift. The enhanced drift

capacity of the shake table test specimen was attributed to redistribution (Moehle and

Diebold, 1984).

Significant yielding of reinforcement in the slabs and columns was measured during the

shake table test. As seen in Table 2-3, equivalent viscous damping ratio measured

during the shake table tests increased rapidly from 1.5% to 7% after the run with 0.189g

peak base acceleration, which induced cracking, yielding, and spalling.

Push-over analyses using ULARC (Sudhaker, 1972), with variable ratios between first

and second floor lateral loading, were conducted to compare with the measured relation

between top drift ratio and base shear. The strength and stiffness of the slabs were

modeled by adopting the effective beam width model and the moment-rotation relations

24
at slab - column connections were modeled using results obtained from tests conducted

by Zee and Moehle (1984). Strength and stiffness of the columns were determined by

analyzing column sections using the modified Kent and Park model (Park et al., 1982)

for unconfined and confined concrete. Figure 2-15 reveals that the analytical model with

2:1 distribution of lateral loads applied at the second and first floor levels, respectively,

compared well with the envelop of the measured responses up to 2% drift. Beyond 2%

lateral drift, a lateral load distribution ratio of 0.5:1 resulted in better predictions. The

unbalanced moment capacity at an interior connection was measured to evaluate the

eccentric shear stress model in ACI 318-83 (“Building”, 1983) and the calculated

capacity of unbalanced moment transfer was 81% of the measured capacity.

Hayes, Foutch, and Wood (1999) conducted a shake table test of a one-third scale,

three-story, 2 × 2 bay lightly reinforced concrete flat plate structure with a gravity shear

ratio Vg/Vo of 0.25 for an interior connection (Fig. 2-16). The objectives of the tests

were to assess the benefits of using viscoelastic dampers as a rehabilitation measure for

low-rise flat plate systems located in moderate seismic zones (UBC-55 Zone III). The

initial seismic simulations were conducted with viscoelastic dampers installed, followed

by the simulations without the viscoelastic dampers. The tests revealed that the

viscoelastic dampers were capable of dissipating more than 90% of the input energy

from the table motion, leading to acceptable performance for the table motions studied.

Figure 2-17 presents the interstory drift response of the second story for two different

ground motions for the simulations without viscoelastic dampers. The plots indicate a

period shift at approximately 4.5 s, when damage to the spandrel beam was observed.

While no serious damage occurred in the connection regions, significant torsional

25
damage occurred in the spandrel beam. Damage to the columns with insufficient lap

spice lengths and wide tie spacing was not observed, apparently due to the confining

effects of the collars that were used to attach the viscoelastic dampers (Fig. 2-16).

2-7. Summary

The effective beam width model and the equivalent frame model have been used to

represent the lateral-load stiffness of slab - column frames. The α-value for the effective

beam width model depends primarily on the column and slab aspect ratios. In the

equivalent frame model, the slab - column frame is represented by using slab-beams

equal to the full span in the direction of loading, as well as a transverse torsion member.

For both lateral-load stiffness models, the effect of cracking on stiffness must be

considered, typically using a β-factor. Several suggestions for appropriate stiffness

reduction factors have been reported, including the recommendations by Hwang (1989).

In flat plate floor systems, shear and unbalanced moment from the slab are transferred

to the column by: (1) direct shear and eccentric shear acting on a critical section, and (2)

flexural yielding over a slab transfer width. Punching shear failure due to the combined

stresses due to unbalanced moment transferred by eccentric shear and direct shear due

to gravity loads is a concern. The eccentric shear stress model in ACI 318-02

(“Building”, 2002) is commonly used to compute the shear strength of slab - column

connections. If the connection strength is insufficient to develop a flexural yield line

across the full width of the slab, punching shear failure will occur prior to formation of

the yield line.

Design requirements for post-tensioned flat plates and flat plates with shear

26
reinforcement are described in ACI 318-02 (“Building”, 2002). Post-tensioned or shear-

reinforced slab - column connections improve the lateral-load stiffness, strength, and

ductility relative to “conventionally-reinforced” slab-column connections; however,

limited test results are available despite of the extensive use of these systems.

Gravity shear ratio has been shown to influence the drift capacity of slab - column

connections. ACI 318-02 (“Building”, 2002) limits the gravity shear ratio Vg/φVc to 0.4

to avoid punching failures which tend to occur for lateral drift ratios exceeding 1.5%. In

addition, continuous bottom reinforcement is required to suspend the slab after

punching failure.

Previous experimental studies of post-tensioned flat plates have led to the following

conclusions: (1) Placing bonded top reinforcement within a distance of 1.5 times the

slab thickness from the face of columns reduces cracking, (2) Providing balanced

quantities of tendons and bonded reinforcement in the connection region improves

hysteretic damping and ductility. (3) Tendon stresses at the ultimate load capacity of the

specimens tested were less than those predicted by ACI 318-77, Eq. 11-13 (“Building”,

1977). (4) A banded tendon arrangement enhances the stiffness and strength of the edge

connections due to high local pressure, and also results in greater displacement ductility.

Shake table tests of flat plate systems with spandrel beams were conducted by Moehle

and Diebold (1984), and Hayes, Foutch, and Wood (1999). Gravity shear ratios Vg/Vo

for the tests were of 0.2 and 0.25, respectively. Punching failures occurred at the first

floor interior connection for the tests conducted Moehle and Diebold (1984), whereas

the tests conducted by Hayes, Foutch, and Wood (1999) exhibited no serious damage in

the connection regions, although significant torsional damage in the spandrel beam was

27
observed.

28
3. Specimen Design and Construction
Two test specimens were constructed at the UC Berkeley Richmond Field Station and

tested on the shake table at the UC Berkeley Earthquake Engineering Research Center.

The specimens represented a reinforced-concrete (RC) flat-plate system and a post-

tensioned (PT) flat-plate system, respectively. Each specimen consisted of a 2 × 2 bay

slab - column frame, two stories high (Fig. 3-1). Due to dimensional limitations of the

shake table, a scale factor of approximately one-third was used for both specimens

based on the typical span-to-depth ratios used for each type of construction (RC and PT).

Each of the specimens is described in detail in the following subsections.

3-1. RC Specimen

Two reinforced concrete prototype buildings were selected to assist in realistic specimen

proportions and detailing (Fig. 3-2). The prototype buildings consisted of a five-story

flat-plate floor system with shear walls designed according to UBC-97 requirements for

zone 4 (high seismic region) and a three-story flat-plate floor system designed for UBC-

97 requirements for zone 2 (moderate seismic region). The span-to-depth ratio of 23

was chosen to satisfy the typical span-to-depth ratios of 20 to 30 used for reinforced-

concrete flat-plate systems on the west coast of the US.

The RC specimen included six columns, with two full spans in the east-west (E-W)

direction, and one full span and two half-spans in the north-south (N-S) direction (Fig.

3-2a). This geometry was selected to provide symmetry and stability, as well as

appropriate boundary conditions for uni-axial shake table tests. Bay widths of 6 ft. 9 in.

(2.06 m), a slab thickness of 3.5 in. (89 mm), and story height of 39 in. (0.99 m) were

29
selected based on the approximately one-third scale factor. Elevation views of RC

specimen are shown in Fig. 3-3.

A yield line solution was employed to assess requirements for column to slab flexural

strength ratios. The yield lines were assumed to extend along the full width of the slab

in either positive or negative bending as required to produce a collapse mechanism (e.g.,

see Fig. 2-3). Column cross sections were selected to be 8 in.× 8 in. (203 mm × 203

mm) reinforced with 8 - # 4 (db=0.5 in.; 12.7 mm) bars with a yield stress of 71.2 ksi

(491 MPa) (Fig. 3-4, Table 2-1). All columns were provided with # 3 (db=0.375 in.; 9.5

mm) closed-hoops spaced at 2 in. (50.8 mm) on center (corresponding to the maximum

spacing of 6 in. or 152 mm allowed for the prototype column) per ACI 318-02, Sec 21.4

(“Building”, 2002). The column hoops were selected to provide sufficient shear strength

to resist the column shear developed for a sway mechanism. No column splices were

used. The column yield moment was 330 in.-k (37 kN·m) for an axial force of 9 kips

(35.5 kN), which was estimated as the axial load due to slab self-weight for the second

story interior columns. For the second floor level, the column yield moment was

sufficient to resist the nominal negative moment that could develop across the full slab

width for an exterior connection (215 in.-k; 24 kN·m), and the sum of the nominal

negative (215 in.-k; 24 kN·m) and positive (160 in.-k; 18 kN·m) moments that could

develop within the column strip for an interior connection. Columns at the first floor

level were sufficient to resist nominal moments for slab yield lines across the full slab

width, as the column extended both above and below the slab.

To monitor base shear and overturning moment, and the unbalanced moment transferred

at each of the first-story slab - column connections during testing, individual footings

30
supported on load cells were used for each column. Eight, one-inch inside diameter

(25.4 mm I.D) conduits were provided in each footing to allow for attachment of the

specimen to the tri-axial load-cells as shown in Fig. 3-5. The depth of the footings was

selected to exceed the development length ldh of the column longitudinal bars. Footings

were well-confined, with 5 - # 4 (db=0.5 in.; 12.7 mm) horizontal and 6 - # 4 (db=0.5 in.;

12.7 mm) vertical stirrups provided (Fig. 3-5).

Slab moments due to combined gravity and lateral load were determined using the

equivalent frame method as described in ACI 318-02, Sec 13.7 (“Building”, 2002). Slab

flexural reinforcement was designed to satisfy ACI 318-02 (“Building”, 2002) and UBC

97 code requirements for nominal flexural strength to resist the slab moments due to

gravity load. Based on requirements of ACI 318-02, Sec 21.12 (“Building”, 2002),

sufficient flexural reinforcement was also placed within the column strip to resist all the

unbalanced moment due to earthquake, as well as tributary column strip gravity

moments. The slab reinforcement consisted of # 3 (db=0.375 in.; 9.5 mm) deformed bars

with a yield stress of 66.4 ksi (458 MPa) having 180-degree standard hooks at the slab

edge (Fig. 3-6a). Concrete clear cover was 3/8 in. (9.5 mm) for both top and bottom slab

reinforcement.

Slab flexural reinforcement placed within c2 + 3h provided sufficient capacity to resist

the fraction of the unbalanced moment transferred in flexure γfMunb, and 64% of the

reinforcement in the column strip were placed within c2 + 3h for an interior or edge

connection (Fig. 3-7, 3-8). The top and bottom reinforcement ratios within c2 + 3h of

1.38% and 0.78%, respectively, did not exceed 0.75ρb = 2.14%, where ρb is the balanced

steel ratio, where c2 is the column dimension perpendicular to the direction of the

31
applied loads and h is the slab thickness. In addition, ACI 318-02 requirements of Sec.

13.5.3 (“Building”, 2002) were satisfied, as 71% of top bars in column strips, and all

bottom bars, were continuous (Fig. 3-6, 3-9c). Table 3-2 summarizes the reinforcing

steel ratios for various conditions.

Given the objectives of the project, particular attention was paid to the design of the

connection regions (Fig. 3-9). Use of stud-rails was chosen to enhance the shear

capacity of slab-column connections, since they are very easy to place and available test

results have shown them to be effective. Design shear strength of the connections was

based on ACI 318-02 (“Building”, 2002) and ACI 421.1R-99 requirements. Stud-rails

were provided to resist combined direct gravity shear stress vdirect and eccentric shear

stress vunb due to the unbalanced moment on the critical section located at d/2 from the

column face (Fig. 3-10a), using (2-16). Outside the shear-reinforced region (Fig. 3-10a),

shear resistance was provided by concrete as given in (2-15).

Stud-rails for the RC specimen consisted of seven studs with top anchor heads welded

to a bottom rail. Detailed information on the studs is provided in Fig. 3-10(b). The shaft

of the stud had a yield stress of 92.1 ksi (635 MPa). The number of studs per rail was

selected to ensure that failure would occur within the shear-reinforced zone, versus

outside the shear-reinforced zone (Fig. 3-10a). Based on ACI 421.1R-99, the first studs

were located at 0.75 in (19.05 mm) away from the column edges, and studs were spaced

at 0.5d = 1.5 in (38.1mm) to intersect the shear cracks properly. Two stud-rails were

placed perpendicular to each column face, or 7.375 in. (187 mm) apart. The spacing

between stud-rails slightly exceeds the manufacturer’s recommended maximum

distance of twice the slab effective depth, or 2d = 2(3 in.) = 6.0 in. (177.8 mm).

32
Normal weight concrete mixed with Type II cement and maximum aggregate size of 3/8

in. (9.5 mm) were used. Two placements were required to construct the RC specimen.

Concrete for the column footings, first-story columns, and first-story slab was placed on

February 15, 2002. Concrete for the second-story columns and second-story slab was

placed on March 11, 2002. Measured mean compressive strength of the concrete used in

the first and second concrete placements were 3870 psi (26.7 Mpa) and 2820 psi (19.4

MPa), respectively, based on core sample tests in accordance with ASTM C-42 (Table

3-1). After finishing concrete surfaces, membrane-forming curing compounds were

sprayed uniformly on top surfaces of slabs to prevent moisture loss. First and second-

story wood forms were removed at the same time, hence the first and second-story

concrete were exposed in air after 41 days and 16 days, respectively.

Concrete compression tests of 6 in. × 12 in. (152 mm × 305 mm) cylinders were

conducted to determine the concrete strengths during the curing period. Table 3-3 shows

the compressive strengths of the first and second placements. For the first batch, the 28-

day strength is 2933 psi (202 MPa), which is less than the target strength of 4000 psi

(27.6 MPa). The concrete strength results from the 6 in. × 12 in. (152 mm × 305 mm)

cylinders for the second placement are not compatible with the concrete strength results

obtained with the core sample tests. It is believed that water was added after acquiring

the concrete used for the cylinders. The concrete cylinders set aside for evaluation on

the day that the specimens were tested on the shake table were misplaced by the

commercial testing laboratory. One remaining concrete cylinder of the first batch was

tested to obtain stress-strain relations at UCLA structural laboratory four months after

the completion of testing of the PT specimen. A peak stress of 3773 psi (26 MPa) and

strain at peak stress of 0.002395 were obtained. The secant modulus of elasticity to

33
0.45fc’ obtained from the stress – strain relation was approximately 2800 ksi (19305

MPa) (Fig. 3-11).

Material properties obtained for the deformed reinforcing bars, stud-rail round bars, and

seven-wire strands are shown in Table 3-4 and Appendix A. Tensile testing was

conducted in accordance with ASTM A370 and Annex A7 of ASTM A370. For stress-

strain relations not displaying a distinct yield point, a strain offset of 0.2 percent was

used to define the yield stress.

3-2. PT Specimen

The configuration of PT specimen was identical to RC specimen (Fig. 3-12, 3-13),

except for the slab thickness and the span length. A slab thickness of 3 in. (76.2 mm)

and bay widths of 9 ft 9 in. (2.97 m) in the direction of testing were selected for the PT

specimen. The resulting span-to-depth ratio of 37.3 is within the range of 36 to 45

typically used for post-tensioned flat-plate systems on the west coast of the US. The

specimens were constructed outside the laboratory; therefore, the width of the

laboratory door (19 ft.; 5.8 m) limited the span length to 8 ft. 8 in. (2.64 m) between

columns transverse to the direction of loading, with overhanging “one-half” span

lengths shortened to 3 ft. 10 in. (1.17 m) on each side (Fig. 3-13b, 3-14). Columns and

footings of the PT specimen were the same as those of the RC specimen, resulting in

smaller column size-to-span length ratio of 7% (c/l = 0.07). The use of identical footing

and column size allowed the reuse of many items.

Slabs were post-tensioned with 5/16 in. (7.94 mm) diameter, seven-wire strand with

yield strength of 235.7 ksi (1625 MPa). The post-tensioning strands were greased and

34
wrapped directly in an extruded plastic coating per PTI recommendations (Post-

tensioning Institute, 2000). A special sized die for the 5/16 in. (7.94 mm) diameter

strand was fabricated for this process, which eliminated the need to push the greased

strands through tubes. As a consequence, the diameter of the unbonded extruded

tendons was kept to 7/16 in. (11 mm), which was desirable given the slab thickness (3

in.; 76.2 mm). Because an off-the-shelf grip for use with 5/16 in. (7.94 mm) diameter

strands was unavailable, a special jack gripper suitable for a Velzy hydraulic jack was

also fabricated.

Tendon arrangements are shown in Fig. 3-14(a) and 3-15(a). Six banded tendons per

span were used in the E-W direction (direction of testing), and seven distributed tendons

per span were used in the N-S direction. Tendons were placed according to ACI 318-99

(“Building”, 1999) and ACI 423.3R-96 provisions and recommendations, respectively.

At least two tendons were placed through the column cages, and the maximum and

minimum distance between tendons was selected such that 0.35 in2 (22.6 mm2) of strand

per bay was provided in the direction of testing. An effective tension force of 9.9 kips

(44 kN) per strand was applied based on flexural strength requirements, resulting in an

average compression stress of 202 psi (1.39 kPa) in the banded (testing) direction (E-W)

and 199 psi (1.37 kPa) in the other direction (N-S). Therefore, the post-tensioned slab

balanced a load of 98 psf (97 MPa), or 85% of dead load (including lead-weights, which

are discussed in more detail in Chapter 4, Section 4.1.).

Anchor plates (3 × 3 × 1/2 in.; 76.2 × 76.2 × 12.7 mm) with 5/8 in. (15.9 mm) diameter

holes were used to provide bearing against the concrete for the tendons (Fig. 3-16a). In

order to transfer the prestressing force to the concrete, reusable barrel anchors were

35
installed at jacking ends and dead ends (Fig. 3-16b). The jacking ends and dead ends

were alternated.

Bonded reinforcement was provided at connection regions to control cracking as

required by ACI 318-02 Section 18.9 (“Building”, 2002). For interior connections, 6 - #

2 (db = 0.25 in.; 6.35 mm) top and bottom bars with a yield stress of 69.8 ksi (481 MPa)

were provided in both directions (Fig. 3-15b, 3-17a). For edge connections, 8 # 2 (db =

0.25 in.; 6.35 mm) hairpin bars, and 4 # 2 (db = 0.25 in.; 6.35 mm) top and bottom bars

were placed in the E-W direction (direction of testing) and N-S direction, respectively

(Fig. 3-15c, 3-17b). Bonded bars provided to control cracking at edge connections also

provided confinement to the anchorage zones (Fig. 3-15c, 3-17b). In addition, edge

tension reinforcement was provided at the anchorage zones to resist bursting pressures

(Fig. 3-16a, 3-17b).

Connection regions of the PT specimen also were reinforced with stud-rails (Fig. 3-15,

3-17) based on requirements of ACI 318-02 (“Building”, 2002) and ACI Committee

Report 421.1R-99. The size and spacing of the stud-rails was selected to require flexural

yielding of bonded bars prior to punching failure on a critical section d/2 from the

column face. The stud-rails used for the PT and RC specimens were identical in

dimension, spacing, and material, except for the length of studs, the number of studs per

rail, and the number of stud-rails per column face. The length of studs depends on the

slab thickness (ACI 421.1R-99, Chapter 6), and 5 studs per rail were chosen to achieve

a conservative shear strength at the outer critical section (and thus ensure failure at the

critical section d/2 from the column face). Three stud-rails were placed perpendicular to

each column face, because the slab thickness (3.0 in.; 76.2 mm) dictates that the spacing

36
between stud-rails not exceed 2d = 4.5 in. (15.24 cm) based on recommendations from

the manufacturer (Decon, 1996).

The PT specimen was constructed at the same time as RC specimen. Sufficient

platforms and shores were provided to support the concrete self-weight of slabs (37.5

psf (37.2 MPa)). Holes were drilled in slab edge forms to allow for the tendons to be

placed (Fig. 3-15). Anchor plates with two - drilled and tapped holes for 1/4’’φ × 20

(6.35 mm diameter) machine threaded fasteners were attached against the inside edge of

the forms, ensuring the plates are perpendicular to the forms (Fig.3-16a, 3-17b). Rebar

chairs were purchased or fabricated to ensure tendons were installed as specified on the

drawings (Fig. 3-18) within a tolerance of 1/8 in. (3.2 mm). Accurate installation of

tendons ensured that wobble friction caused by unintended curvature was minimized.

Clear cover was 3/8 in. (9.5 mm) over both the top and bottom of strands or

reinforcement except at interior connections where top clear cover of 5/16 in. (7.9 mm)

and 1/2 in. (12.7 mm) were designated for tendons and deformed rebar at an interior

connection region, respectively due to space limitations (Fig. 3-18a).

The PT specimen was cast in place using the same concrete as was used for the RC

specimen. After casting, concrete slabs were sprayed of curing compounds. Mean

compressive strengths of 2730 psi (18.3 MPa) and 2990 psi (20.6 MPa) were obtained

for the first-story and second-story slabs of PT specimen, respectively, based on core

sample tests done in accordance with ASTM C-42 (Table 3-1). All wood forms for the

PT specimen were removed at the same time as the removal of forms for the RC

specimen, except for the slab edge forms, which were detached just prior to post-

tensioning.

37
Before shores were removed, all tendons were partially tensioned to an average of 4.4

kips (19.6 kN) per tendon to balance the slab self-weight. The jack pressure was

calibrated by using donut-shaped load cells, which measured the prestressing force

directly (Fig. 3-19). Partial post-tensioning work for the first and second stories was

conducted 5 and 8 days after each concrete placement, respectively. The post-tensioning

process first involved removing the slab edge forms and removing all exposed plastic

sheathings on the 5/16 in. (7.94 mm) strands beyond the bearing plates. Protruding

strands were cut-off 17 in. (43 cm) from the slab edge prior to moving the specimen into

the laboratory. The 17 in. (43 cm) length was sufficient to allow the post-tensioning jack

to grip the tendons for final post tensioning, which was done during the installation of

the lead weights. The installation of the lead weights and the final post-tensioning are

discussed in Chapter 4, Section 4.1.

38
4. Instrumentation and Experimentation
After construction was completed, the specimens were moved into the laboratory and

placed on the 20 by 20 ft. (6.09 m x 6.09 m) earthquake simulator. During testing,

information was collected from 193 channels for each specimen. This chapter describes

the setup and instrumentation of the specimens, as well as details associated with the

data acquisition process. In specific, details are provided on: (1) moving the specimens,

(2) attaching the specimens to the table, (3) installing the lead-weights on the floor slabs,

and (4) attaching the various types of instrumentation to the specimens. Information on

the table acceleration histories applied to the specimens also is provided.

4-1. Test Setup

The specimens were braced extensively prior to moving them into the laboratory

building using rollers and a forklift. The bracing system consisted of L2.5×2.5×0.25

(63.5 mm × 63.5 mm × 6.35 mm) angles for diagonal braces and C12×20.7 (30.5 cm ×

302 N/m) base horizontal channels as shown in Fig. 4-1. The bracing systems were

conservatively designed based on results of an elastic analysis using SAP 2000

(Computers and Structures, Inc., Nonlinear Version 7.42) to avoid concrete cracking

during the moving process. Prior to casting the concrete, horizontal treaded rods were

embedded within the columns and footings to attach the bracing elements (Fig. 4-2). At

bolted connections, B-11 grout was placed between the concrete and brace surfaces to

minimize any unintentional flexibility of the bracing system. After moving the

specimens into the laboratory, the specimens were painted with white latex paint and 4

in. (10.2 cm) square grids were drawn at connection regions to provide a visual

39
reference system for cracks observed during the tests.

Once inside the laboratory, the RC specimen was moved onto the shake table using a

10-ton bridge crane. Prior to concrete casting, an anchor was inserted at the top of each

column to allow for lifting (Fig. 4-3). Since the weight of PT specimen exceeded the

crane capacity, the PT specimen was lifted and moved onto the shake table by using a

forklift on one side and the crane on the other side of the specimen.

A pair of 20 ft x 3 ft (6.1 m x 0.9 m) long, 2 in. (50.8 mm) thick base steel plates were

used to attach the RC and PT specimens to the shake table as shown in Fig. 4-4. Based

on the geometry of each specimen, holes for 7/8 in. (22.2 mm) diameter bolts were

drilled and tapped into the base plates (Fig. 4-5), allowing the baseplate to be used for

both specimens without removal and realignment. A layer of hydrostone grout was

placed between the base plates and the table to ensure a level surface, and the plates

were prestressed to the shake table platform using the existing holes spaced at 3 ft (91.4

cm) on center in the shake table platform. Additional anchorage plates were added

because the centerline of the footings of the specimens did not align with the holes in

the shake table platform (Fig. 4-4).

Tri-axial load cells were bolted to the base plates using four, 4 in. (101.6 mm) long,

Grade 8, 7/8 in. (22.2 mm) diameter bolts with 2 in. (50.8 mm) of thread. A 16 in. × 16

in. × 2 in. (40.6 cm × 40.6 cm × 50.8 mm) steel plate was used to connect the load cells

to the footings as shown on Fig. 4-6. The footings for the six columns were not tied

together (i.e., each column was supported on an isolated footing) so that the readings

from the load cells at the base of each footing could be used to obtain footing and

column base reactions. The tri-axial load cells were levelled and grouted at the top and

40
bottom. Afterwards, specimens were seated on top of the footing steel plates, which

were placed on top of the load cells. Eight of Grade 8, 3/4 in. (19.05 mm) diameter

threaded rods were inserted through the conduits embedded in the footings and bolted to

the 16 in. × 16 in. × 2 in. (40.6 cm × 40.6 cm × 50.8 mm) steel plate. Three-quarter inch

(19.05 mm) thick steel washers were placed on the top and bottom of the threaded rods

through the footings to avoid localized concrete crushing (Fig. 4-7).

Given the approximate 1/3 scale factor on geometry, ground motions were time-

compressed by (1/3)0.5 and forces were scaled by (1/3)2. Typical ready mix concrete and

steel reinforcing bars were used; therefore, stresses or pressures were not scaled. To

account for the higher material capacities, lead weights were added to the floor slabs

after the specimens were mounted and levelled on the shake table to simulate the inertial

and gravity stresses of the prototype building (Fig. 4-8) based on artificial mass

simulation (Harris et al., 1999). The lead weights provided 23.2 kips (103.2 kN) of

superimposed weight per floor for the RC specimen, resulting in total uniform weight of

165 psf (7.9 kPa) at each floor level. The 165 psf (7.9 kPa) uniform floor load

represents the self-weight of the prototype floor slab of 10.5 in. (266.7 mm) thickness,

or 135 psf (6.46 kPa), as well as an additional dead load of 30 psf (1.35 kPa). The total

slab weight at each level of the PT specimen (specimen and lead weights) is 24 kips

(106.8 kN), or approximately 115 psf (5.51 kPa), which corresponds closely to the self-

weight of the prototype post-tensioned slabs of 9.0 in. (228.6 mm) thickness, or 115 psf

(5.51 kPa).

As a consequence of scaling relation and the added lead weights, the gravity shear ratios

for the interior connections were 0.26 and 0.30 for the RC and PT specimens,

41
respectively, provided that the design concrete strength (fc’) is 4.0 ksi (27.6 MPa). The

gravity shear ratios are one of most important factors affecting the behavior of flat

plates (Pan and Moehle, 1989). The gravity shear ratios of 0.26 and 0.30 selected for the

test specimens are close to those used in typical construction.

The lead weights used were single ingots, each weighing 100 lbs (445 N) and with a

footprint of 4.375 in. × 21 in. (11.1 cm × 53.3 cm) and a height is 3.5 in. (89 mm). A

majority of the lead weights positioned on the specimens were bolted in bundles of five

(Fig. 4-8, 4-9). A total of 80 and 96 bundles were used for RC and PT specimens,

respectively. An additional 64 ingots, stacked two high, were placed on the RC

specimen (Fig. 4-9). An equivalent frame analysis was conducted for each specimen to

assist in determining the positioning of the lead weights to reproduce the slab moments

and shears for gravity load only. The crane was used to position lead weights on the top

floor slab; however, it could not be used for the first floor slab. A pair of steel rollers

were developed and fabricated to skate the heavy lead weights to the designated

locations from the edge of the first floor as shown in Fig. 4-10.

The lead weights were held in place with two steel angles and four 3/8 in. (9.5 mm)

diameter threaded rods with bolts and washers (Fig. 4-9). The weights were isolated

from the slab using steel and rubber pads based on findings reported by Moehle and

Diebold (1984). Steel pads were used at one side to provide rigidity needed to have the

lead weights act as floor mass during the tests, whereas rubber pads were used at the

other side so that the weights would not stiffen or strengthen the slab (Fig. 4-11). For

bundles of five ingots, two steel pads were placed on the same side of the bundles in the

direction of testing (E-W direction), to avoid stiffening or strengthening the slab in the

42
direction of shaking. The lead weights were anchored to the slab using threaded rods

(Fig. 4-9). Three-eighths inch (9.5 mm) diameter wooden dowels were attached to the

forms and oiled prior to concrete casting to provide the anchor points for the lead

weights (Fig. 4-12). The dowels were punched out to provide the holes through the slab

needed to anchor the lead weights.

To minimize potential for slab cracking during the installation of the lead weights, the

first floor slab of PT specimen was re-shored with 24, 4 in. × 4 in. (10.2 cm × 10.2 cm)

posts and wooden shims prior to installation of the lead weights (Fig. 4-13). Re-shoring

was necessary due to the difficult installation process for the lead weights on the first

floor slab. For the second story slab, placement of lead weights was easily controlled.

Therefore, only a portion of the lead weights were initially placed on the slab, followed

by increasing the post-tensioning force for the first and second floor slabs from the

force required to support slab selfweight, to the force required to resist slab selfweight

and the total weight of the lead at each floor level. The remaining lead weights were

then added to the second floor slab and the shores for first floor slab were removed.

4-2. Instrumentation

An instrumentation scheme was devised to acquire the data to assess important

modelling and design issues, such as: shear and moment transfer at connections, force

and moment redistribution, stud-rail forces and deformations, post-tensioning forces,

and dynamic responses. A total of 193 data channels were collected at 100 Hz (0.01 sec).

Data from all channels were filtered to remove frequency content greater than 100 Hz.

A summary of the data collected is provided in Tables 4-1 and 4-2, with additional

details provided in the following subsections.

43
4-2-1. Table Instrumentation

A total of 16 channels of data were dedicated to table motions, including: (1) 8

accelerometers attached under the table for measuring three dimensional table

accelerations, as well as table pitch and roll accelerations, and (2) 8 displacement

gauges for monitoring the displacement of the table actuators.

4-2-2. Strain Gauges

Strain gauges were attached on selected column and slab reinforcing bars, as well as

stud-rails (Fig. 4-14). Gauges for #3 (db = 0.375 in.; 9.5 mm) bars and #2 (db = 0.25 in.;

6.35 mm) were 5 mm and 2 mm long, respectively. Model YFLA gauges from Texas

Measurements, with a maximum strain of 20 %, were used for the #2 (db = 0.25 in.;

6.35 mm) and #3 (db = 0.375 in.; 9.5 mm) reinforcing bars. Strain gauges were attached

to the reinforcement prior to placing the reinforcement in the forms. Strain gauges

attached on column reinforcing bars were located at approximately 1 in. (25.4 mm)

from the column – slab interface, whereas gauges attached on slab reinforcement were

located approximately 1 in. (25.4 mm) away from the column on a line extending across

the slab as shown on Fig. 4-15. Strain gauges monitored strains on reinforcing bars and

stud rails, and pairs of gauges were used to monitor column and slab curvatures. The

selected reinforcing bars on which strain gauges were attached are shown in Fig. 4-16

(RC specimen, top bars), 4-17 (RC specimen, bottom bars), 4-18 (RC specimen, column

bars and stud-rails), 4-19 (PT specimen, top and bottom slab bars), and 4-20 (PT

specimen, column bars and stud-rails). Concrete surface strain gauges used for the PT

specimen test are shown on Fig. 4-19 and 4-20.

44
Very careful attention was paid to the installation of the strain gauges. First, the rebars

were filed at designated locations and abraded with acetones, conditioners (MCA-2;

Micro Measurements) and neutralizers (MN5A-1; Micro Measurements) to make

smooth and clean surfaces; however, no more than one-half of the bar perimeter was

filed to minimize the loss of bond. The strain gauges and terminals were affixed to the

bars using adhesive (CN-Y; Texas Measurements) and lead wires were soldered to the

strain gauge terminals. The soldered connection was cleaned with rosin solvents (RSK-

1; Micro Measurements) and all surfaces were cleaned with isopropyl alcohols (GC-6;

Micro Measurement) prior to coating the gauge with polyurethane (M-coat A; Micro

Measurements), which was allowed to cure for one day. The strain gauges were also

waterproofed with an epoxy coating of nitrile rubber (M-coat B; Micro Measurements),

which was allowed to cure for two hours prior to applying a compound of mixed

polysulfide and polymer (M-coat J-3; Micro Measurements). The compound was

applied only over the gauge, and did not extend around the bar, to preserve bond along

the remaining portion of the bar perimeter (Fig. 4-14b).

After testing of the RC specimen, there were concerns that the displacement gauges did

not capture very small deformations effectively; therefore, 120 mm long polyester wire

strain gauges (P Series, Texas Measurements) were mounted on the surface of some

columns of the PT specimen (Fig. 4-19, 4-20).

Strain gauge readings were recorded before and during the application of the lead

weights.

4-2-3. Load Cells

45
A total of six tri-axial load cells were installed, one under each column footing at the

base of the specimen (Fig. 4-6). A schematic of a tri-axial load cell is shown in Fig. 4-

21. The load cells were made from heat-treated ASTM 4140 steel to obtain maximum

hardness, resulting in a proportional limit of 130 ksi (896 MPa). Calibration of the load

cells has revealed that results are sensitive to end conditions; therefore, careful attention

was paid to properly grouting the end plates supporting the load cells to the steel plate

attached to the shake table and the plate connecting the load cell to the column footing.

The tri-axial load cells were used to measure shears and overturning moments at the

base of the footings in two directions, as well as vertical (axial) forces. Twisting forces

was not obtained.

Four 15 kip (66.7 kN) maximum capacity, donut-shaped load cells (Model D, Sensotec)

were installed to measure the strand forces during post-tensioning, as well as changes in

strand forces during testing. The post-tensioning strand passed through a 1/2 in. (12.7

mm) hole in the load cell. The load cell was placed between two flat and parallel anchor

plates as shown in Fig. 4-22 and 4-23. Two load cells were positioned on a post-

tensioning strand in the line of loading at both the jacking end and the dead end, and the

other two load cells were positioned on two strands in the line of loading at the jacking

end. All of the strands monitored with load cells were located on the first floor slab in

the north frame (Fig. 4-22).

4-2-4. Accelerometers

Accelerometers were mounted on top of the table and on the top of each floor slab (Fig.

4-24). Since accelerations measured on the underside of the shake table may not

46
correspond to the table accelerations on the top side of the shake table, two

accelerometers were mounted on top of the table to measure transverse accelerations in

the testing direction, and transverse to the testing direction (Fig. 4-25). An aluminium

block, on which two accelerometers were attached, was anchored to the center of the

table with epoxy (Fig. 4-24). Four accelerometers were attached at each floor level of

the specimens to monitor transverse accelerations in two directions. Accelerometer

locations for the RC specimen are given on Fig. 4-26 and 4-27, and locations for the PT

specimen are given on Fig. 4-28 and 4-29. Accelerometers attached to the top of the

table and to the floor slabs were piezoresistive OEM accelerometers with +/− 5g range

(Model 302, EG&G IC Sensors), whereas linear accelerometers that produce high level

DC outputs were used for shake table underside (Model 141, Setra).

4-2-5. Displacement Gauges

Lateral displacements at each floor level and at the top of the footing were measured

using displacement transducers (DCDTs) and wire potentiometers (WPs; Model PT-101,

Celesco). A DCDT (displacement transducer DC-DC series 240, Trans-Tek) is linear

variable differential transformer (LVDT) adopting DC input and DC output. When the

core moves within the coil assembly, an oscillator and a demodulator produce DC

voltage changes proportional to the core displacement. A core with a diameter of 0.099

in. (2.5 mm) was chosen to reduce the friction within a 0.125 in. (3.18 mm) inner

diameter hole. The DCDTs were used because they provided the high resolution

measurements needed to obtain reliable readings at low and moderate intensity shaking

levels. Wire potentiometers were used to extend the displacement range over which data

could be collected for the moderate-to-high intensity shaking, as discussed in the

47
following paragraph. The locations of DCDTs and WPs used to monitor lateral

displacements are shown in Fig. 4-30 and 4-31 for the RC specimen and Fig. 4-32 and

4-33 for the PT specimen.

Absolute displacements were obtained by mounting the displacement transducers and

potentiometers between the specimen and rigid external reference frame located

adjacent the shake table. Relative displacements between floor levels were obtained by

computing the difference in absolute measurements. The DCDTs used had a linear range

of 3 in. (76.2 mm) and monitored absolute displacements during the shake table and the

free vibration tests. For the high intensity shake table motions (generally Runs 4 and 5),

the absolute displacements exceeded the linear range of the DCDTs; therefore, wire

potentiometers were mounted in parallel with the DCDTs to obtain the displacement

readings where the linear range of the DCDTs was exceeded. The linear range of the

wire potentiometers is +/− 15 in. (12.7 cm).

DCDTs with linear ranges of 0.5 in. (12.7 mm), 1 in. (25.4 mm), and 2 in. (50.8 mm), as

well as Spring Potentiometers (SP) with linear ranges of 1 in. (25.4 mm) and 2 in. (50.8

mm) were employed to measure average concrete strains on the slabs and columns. The

SPs (TR/TRS Series, Novotechnik Siedle Group) measure displacement using

conductive-plastic resistance. A photo of the attachment at an exterior and interior

column is shown in Fig. 4-34 and 4-35, respectively, with close-up views at the top and

bottom of a column shown in Fig. 4-36 and 4-37, respectively. Threaded rods with 3/8

in. (9.5 mm) and 1/2 in. (12.7 mm) diameter were cast in the slabs and columns,

respectively, to mount the displacement transducers (DCDTs and SPs) used to obtain

average concrete strain (Fig. 4-36). The difference in the measurements from two

48
transducers mounted on either side of a slab (top and bottom; Fig. 4-34) or column (Fig.

4-35) were used to calculate average curvature over the gauge length of the transducers.

Column cover concrete was removed around the threaded rods to avoid local bending of

the treaded rods that might occur due to cracking and spalling of concrete (Fig. 4-36, 4-

37). The circular voids around the treaded rods in the columns were created by

installing waterproofed ethafoams on the inner side of the forms.

The locations of the DCDTs and SPs were determined to acquire the data needed to

assess important modelling and design issues, such as: shear force and unbalanced

moment transferred at the slab-column connections (Fig. 4-25 to 4-30), moments at the

top and bottom of each column (Fig. 4-30, 4-32), and the distribution of slab curvatures

adjacent to the slab-column connection and at slab mid-span (Fig. 4-25 to 4-29). The

locations and types of displacement gauges are summarized in Table 4-1 and 4-2.

Average curvatures were monitored by embedding a pair of threaded rods within the

concrete at critical regions (Fig. 4-36, 4-37). At one rod, a steel washer with a small

hole was fixed with nuts, whereas at the other rod, an aluminium bar and a pulley were

attached. The 6061 T6511 aluminium bars used in conjunction with the DCDTs were

fabricated with a hole drilled through the threaded rod at one end and two small holes at

the other end, and mounted as shown in Fig. 4-36. A hook was fixed at the end of the

rod and connected to a wire which passed over a pulley and connected to the core of the

DCDT. Aluminium bars for SPs were fabricated in a manner similar to bars fabricated

for the DCDTs, except that SPs were attached directly on the bar (Fig. 4-37). An

additional spring was added to the SPs to provide an instantaneous restoring mechanism

for the SP shaft during the dynamic tests.

49
During the tests of the RC specimen, significant footing rotation was observed.

Therefore, prior to testing of the PT specimen, SPs were installed between the footing

and the steel plate used to anchor the specimen to the shake table to monitor the rotation

of the footing relative to the plate (Fig. 4-38). The plane of rotation was monitored for

an interior footing and two exterior footings using three SPs per footing as noted on Fig.

4-39(a). Locations of the SPs relative to the column center are provided in Figure 4-

39(b).

4-3. Testing

Testing was conducted on the earthquake simulator at the UC Berkeley Richmond Field

Station. The specimens were subjected to several runs of uniaxial shaking, with the

intensity of shaking increased for each subsequent run. Preliminary analytical studies

were conducted to assist in the selection of an appropriate ground motion to use for the

testing. The CHY087W motion, recorded on soft soil during the 21 September 1999

earthquake in Taiwan was selected. The CHY087W record is fairly long duration and

does not include significant near-fault features, as the closest distance to fault was 34.46

km (PEER Strong Motion Database). Preliminary analytical studies using DRAIN-2DX

(Prakash et al., 1993) indicated that it would cause sufficient inelastic response to

produce damage to the slab - column connections. The peak values of the spectral

pseudo-acceleration response spectrum were located between 0.2 - 0.4 sec, where the

fundamental periods of the specimens were located. The acceleration history of the

record for each run is shown in Fig. 4-40 and 4-41 for the RC and PT specimens,

respectively. Spectral pseudo-acceleration relations are shown in Fig. 4-42. The

characteristics of ground motion are summarized in Table 4-3.

50
The original ground motion was time-compressed by 1/ 3 corresponding to the

dimension scale factor of 1/3 as shown in Fig. 4-43. The accelerations of the original

ground motion were amplified to provide increasing intensity with each test run. The

tests represented: (1) low-level excitation, (2) service-level excitation to approximately

2/3 of the yield displacement, (3) moderate-intensity excitation to produce limited

yielding, and (4) damage-level excitation. For the PT specimen, a fifth test was run with

very intense motions given that relatively little damage was observed after Run 4. The

ground motion histories and the response spectra for all test runs are shown in Fig. 4-40

to 4-42.

Applied table motions are limited by the peak displacement and velocity of the shake

table, which is 5 in. (12.7 cm) and 25 in./sec (63.5 cm/sec), respectively. Because the

peak table displacement of the original ground motions exceeded the displacement

capacity of the shake table, the ground motion records used for each run were modified

to decrease the maximum displacement by cutting off frequencies less than 0.15 Hz

(For PT run 5, frequencies less than 0.3 Hz were removed). In addition, frequency

content exceeding 15 Hz also was removed.

The plan dimensions of the shake table are 20 ft (6.09 m) × 20 ft (6.09 m) with a

payload capacity of 100 kips (445 kN). The peak table acceleration at payload capacity

is 1.5g. Although the shake table is capable of representing motions for six degrees-of-

freedom at the base of a structure (three translational, three rotational), testing was

conducted for only one horizontal ground motion component. The table driving system

consists of eight, horizontal, 75 kip (334 kN) hydraulic actuators, and four, vertical, 75

kip (334 kN) hydraulic actuators. Swivel connections are used at the actuator ends,

51
allowing rotation without coupling. The primary shake table control system is a MTS

model 469 controller.

Pull-back tests followed by free vibration response tests were conducted between the

shake table runs to assess if the initial stiffness of the specimens had changed, as well as

fundamental periods and damping ratios of the specimens. The average of fundamental

periods and damping ratios are shown in Table 4-4 and 4-5.

52
5. Experimental Results

Observations and preliminary results obtained from evaluation of the data collected

from the shake table tests are presented and discussed. Information presented include:

(1) base shear versus relative displacement relations, (2) column and slab moment

histories, and (3) post-tensioning force histories, as well as (4) observed damage, and

(5) the assessment of slab - column punching failures.

Overall, the test specimens performed well, with the shear reinforcement limiting the

extent of the punching damage compared with tests on specimens without shear

reinforcement. Excellent performance was observed for the post-tensioned slab -

column frame.

5-1. Base Shear versus Relative Displacement

DCDTs and WPs were used to monitor the absolute displacements of each floor slab

level and at the top of the footing. Top displacement relative to the column base was

obtained by subtracting the footing absolute displacement from the second floor

absolute displacement. Since the second floor level displacement includes displacement

due to footing rotation relative to the base steel plate level, corrections were applied to

obtain the interstory displacements (Fig. 5-1).

Average footing rotations were obtained from measurements taken from three SPs on

two exterior and one interior footings during tests of the PT specimen. Averages values

of footing rotation from the interior and the two exterior footings were nearly equal (Fig.

5-2). Footing rotation measurements were not obtained during the tests of the RC

53
specimen; however, the footing rotations that occurred during the tests of RC specimen

were estimated using the data acquired from the tests of the PT specimen. The relation

between footing rotations and moments measured from the tri-axial load cells at the

base of each footing indicated that this relationship was nearly linear; hence, a linear fit

to the moment rotation relation for the PT specimen was determined. The footing

rotation for the RC specimen was determined directly from the measured moment from

the tri-axial load cell at the base of the footing from the tests of the RC specimen, and

the linear fit of the moment versus rotation relation from the test of the PT specimen

(Fig. 5-2).

Base shear versus top relative displacement data were not obtained for Run 1 and 3 for

the RC specimen. The resolution of the WP used to monitor lateral displacements for

Run 1 was insufficient to derive relative displacements. During Run 3 of the RC

specimen, the gain on the load cell moment was not set properly, resulting in clipped

data.

Base shear versus top relative displacement between the column base and the second

story were determined and are plotted in Fig. 5-3 for the RC and PT specimens. Figure

5-3 reveals that the RC specimen was subjected to drift levels of approximately 3% with

only moderate strength deterioration. Deterioration in the lateral load capacity of the

post-tensioned floor system is observed beyond 3% lateral drift. During the first

significant yielding excursion into the nonlinear range, a slight decrease of lateral load

strength of the RC specimen was observed. In contrast, more substantial deterioration in

lateral load capacity was observed for the PT specimen. Results for the RC specimen

reveal significant pinching of the hysteresis loops relative to the PT specimen. The loss

54
of stiffness due to punching of the slab - column connections is apparent for both

specimens. For both positive and negative base shear, lateral load capacity of the RC

and PT specimen deteriorated in subsequent cycles to approximately the same

maximum drift ratio.

The hysteresis loops for PT specimen are narrower than the hysteresis loops for the RC

specimen, indicating that more substantial slab yielding in flexure occurred in the RC

specimen prior to punching failure then for the PT specimen. For the PT specimen, after

yielding of the bonded reinforcement within the slab - column connections occurred, the

PT specimen acts as a semi-elastic structure with very low stiffness due to the presence

of the post-tensioning reinforcement. This behavior is apparent in Run 5, as the stiffness

of the initial cycles of Run 5 match closely the stiffness for cycles at the end of Run 4.

5-2. Column Moment Evaluation

Moment versus curvature relations at the base of one interior and three exterior, first

story columns for the RC and PT specimens are plotted in Fig. 5-4. The column

moments at the base were derived using shear forces and overturning moments

measured from the tri-axial load cells. The load cell measurements were assumed to act

at the centerline of the load cell, and equilibrium requirements were used to determine

the shear and moment at the column base. Column curvatures were calculated using

strain readings from gauges on the column vertical reinforcing bars on opposite faces,

divided by the distance between the gauges (assumed to be the distance between the bar

centerlines). The relations plotted in Fig. 5-4 indicate significant yielding occurred

during the damage level testing, with greater inelastic curvatures produced for negative

loading. Damage at the base of a column of the RC specimen after Run 4 is shown in

55
Figure 5-5.

Evaluation of experimental results to assess whether the data collected are consistent

with expectations, or with other independently obtained measurements, is an important

aspect of experimental work. To assess the consistency of the data collected, three

approaches were used to derive the column moments at two locations (base of a first

story column, and at the top of a first story column). First, column moments were

calculated from the load cell measurements as noted in the preceding paragraph. For the

other two approaches, column moments were obtained using column curvature

estimates derived from experimental data and analytically derived, monotonic, moment

- curvature relations to relate experimentally derived curvatures to column moments.

Column curvatures were calculated as the difference between either two rebar strain

gauge readings or two displacement gauge readings obtained on opposite sides of a

column, divided by the distance between the gauges. Column moments for the two

curvature estimates were determined directly from the computed moment - curvature

relation for each column (Fig. 5-6). The column moment - curvature relations were

computed using the BIAX program (Wallace, 1992), using measured material properties

and accounting for the axial force level measured from the load cells at the bases of the

columns. Although the measured axial force levels from load cells do not appear very

reliable, apparently because the axial load reading is very sensitive to load cell end

conditions (Personal communication with Wes Neighbour, 2002), variation of axial load

within the range observed from the load cell readings does not influence column base

moment significantly.

Column moments for the RC specimen obtained using the three approaches are shown

56
on Figure 5-7 for the first significant excursion into the inelastic range. Moment values

obtained using the three approaches are fairly consistent, with the results from the load

cell and the rebar strain gauges in good agreement.

Because column base moment values are likely to be fairly insensitive to the curvature

reading after yield (e.g., see Fig. 5-6), curvature values are compared directly in Fig. 5-8.

The moments at the first column base from the reading of load cell is linearly converted

to the curvature. The linear relation is derived from the average yield moment (265 in.-

k; 29.9 kN-m) and yield curvature (0.0006/in.; 0.000236/cm) of the first story columns

computed using the BIAX program (Wallace, 1992). The results show that the

curvatures from rebar strain gauges and displacement gauges are very similar. Results

from the load cell compare well prior to yield, which occurs at time equal to 11 seconds

(because the load cell remains elastic, the curvature value never exceeds approximately

-0.001/in. (-0.0039/cm) to 0.0008/in. (0.00315/cm)).

In the preceding paragraph, moment values were compared using three different

approaches to assess the consistency of the results obtained. This exercise was

undertaken because at some locations within the test structures, column moment could

only be obtained using one of the approaches noted, and determination of the column

moment was an essential step in determining the unbalanced moment transferred at the

various slab-column connections. The comparisons indicate that the results obtained by

the three approaches are in reasonable agreement; therefore, column moments at various

locations throughout the specimen were obtained using one or more of the approaches,

depending on the instrumentation available. In general, strain gauge data were used over

displacement gauge data at locations where both data sets existed.

57
The column moments at the first and second floor level slab-column connection regions

did not exceed the yield moment for all tests; therefore, column moments for these

locations were determined using fully cracked flexural stiffness obtained from the

computed moment - curvature relation, multiplied by the experimentally derived

curvature using either rebar strain gauges or displacement gauges. With column

moments evaluated, the next step involved evaluating the moment transferred to the

column by the slabs (interior) or slab (exterior) framing into the column.

5-3. Slab Moment Evaluation

Strain gauges attached to rebars and displacement gauges anchored between two

reference points were used to measure rebar strains and average concrete strains at the

top and bottom regions of the slab, respectively. The rebar strain gauges were located

approximately 1 in. (25.4 mm) from the column face, where rebar yielding was

expected (Fig. 5-9). In contrast, displacement gauges measured average deformations

over a distance equal to approximately one slab thickness (Fig. 5-9) between two

threaded rods located approximately 1.5 in. (38.1 mm) and 5 in. (127 mm) from the

column face, respectively. The consistency of measurements obtained from these

gauges can be assessed using symmetry. For example, Figure 5-10 shows the response

histories of strain gauges located symmetrically on either side of an exterior column.

Results reveal excellent agreement in measured strains for both top and bottom bars.

Similar results were obtained for gauges at other locations.

Rebar strains for top and bottom bars are plotted on Fig. 5-12. Measured strains indicate

that both top and bottom bars within the ACI 318-02 transfer width of c2 + 3h (18.5 in.;

47 cm) yielded. Figure 5-13 also reveals that top and bottom bars within the transfer

58
width of c2 + 3h (18.5 in.; 47 cm) at an interior connection reached yield. The top bar

placed at 5.5 in. (14 cm) from the interior column centerline (Top bar 3) barely reached

yield (~0.002 strains), and two top bars located symmetrically 16 in. (40.6 cm) to the

either side of the column centerline (Top bar 1 and 4) reached approximately 50% and

70% of the yield strains (0.001 and 0.0014) for these bars, respectively. The top bar

located at the center of the span between two interior columns reached approximately

60% of the yield strain (0.0012).

Slab rebar strain and displacement gauges were used to estimate curvature distributions

across the full slab width at the slab-column interface. The curvature distributions were

used to derive slab moments for various widths (e.g., full width, c2 + 3h). Slab curvature

distributions for an exterior connection of the RC specimen near peak displacement

responses for the high-intensity run (Run 4) are shown in Fig. 5-11. The curvature

values obtained using displacement gauges indicate that the reinforcing within the

column strip (40.5 in.; 102.9 cm) and the width of c2 + 3h (18.5 in.; 47 cm) should have

reached yield for both positive and negative curvature. The results from displacement

gauges reveal that the flexural transfer width for interior connections is greater than c2 +

3h, as noted by other researchers (see Chapter 2, Section 2-2).

Crack patterns at exterior connections indicate that the transfer width was limited due to

torsional cracking, as shown on (Fig. 5-14). The approximately 45-degree torsional

“yield lines” are consistent with assumptions used for the ACI 318-02 provisions, where

the transfer width at exterior connections is limited to c2 + 3h or c2 + 2ct. For the PT

specimen, flexural cracks were developed across the full slab width at both exterior and

interior connections as shown in Fig. 5-15.

59
Moments derived for the columns and the slabs can be used to compute the unbalanced

moment transferred at the slab - column connections. Based on the slab curvature

distributions, total slab moment (from slab edge to the centerline of the slab between the

columns) for the various connections at a given time can be computed. The unbalanced

moment transferred at a slab - column connection can be determined by summing the

slab moments. The portion of the slab unbalanced moment transferred in flexure is

assumed to be the portion of the slab moment within the flexural transfer width c + 3h.

The remaining portion of slab unbalanced moment (outside c + 3h) is assumed to be

transferred by eccentric shear. The unbalanced slab moment is resisted by the columns

(1st floor) or column (2nd floor). Column moments determined from strain or

displacement gauges can be compared with the column moments computed based on

slab equilibrium requirements. At this time, this comparison has not been made.

5-4. Observed Damage

At both interior and exterior connections, significant cracking was observed on both the

top and bottom of slabs around the column. Significant torsional cracks were observed

at exterior connections of the RC specimen (Fig. 5-14), as well as at interior

connections (Fig. 5-15). The most significant damage was observed at the exterior

connections of the RC specimen where substantial concrete spalling occurred on the

west edge of the RC specimen during Run 4 (Fig. 5-16). For the PT specimen, cracks

were observed to extend along the full width of the slab (Fig. 5-16), and “torsional”

cracks also were observed at the exterior connections for the PT specimen outside the

region containing the banded post-tensioning (Fig. 5-18). In general, the damage

60
observed at the RC connections (Fig. 5-14, 5-15) was more significant and more widely

distributed than those observed for the PT specimen (Fig. 5-17, 5-18).

Punching of connections occurred during the tests, as will be discussed later, despite the

somewhat limited damage observed within the connection regions compared with

observed punching failures for connections without shear reinforcement (e.g., compare

Fig. 5-14 and 1-4). Significant relative rotation between the slabs and the columns were

observed during the Run 4 for both the RC and PT specimens, indicating that moment

transfer strength had been lost and the connection was free to rotate, like a “hinge”. In

addition, at the completion of Run 4 for each specimen, use of hammers to tap the

concrete around the slab - column connections indicated that the integrity of the slab

concrete had been compromised.

Based on the observed damage, it was concluded that connection punching had

occurred; therefore, data were studied to assess the combination of unbalanced moment,

shear (direct, eccentric), and drift that resulted in connection failure.

5-5. Punching Failure

For isolated connections tested under static, monotonically increasing loads for each

applied load or drift cycle, it is relatively easy to determine when connection punching

occurs, as the lateral-load capacity of the specimen experiences a sudden drop. In

contrast, for dynamic tests of frame systems under simulated earthquake loading, the

specimen undergoes continuous loading and unloading; therefore, a more detailed

evaluation is required to assess when punching failures occur. When punching failure

occurs, the slab loses its ability to transfer moment from the slab to the column;

61
therefore, one approach to assessing when connection punching has occurred is to plot

the relationship between slab curvature and column curvature. As slab curvature

increases, the column curvature should also increase, unless slab moment capacity drops

(i.e., punching occurs) or the column yields. Column yielding is not expected since the

column has been designed to be stronger than the slab; however, if the column were to

yield, the column curvature should increase significantly and the slab curvature should

remain relatively constant or increase slightly.

Slab curvature versus column curvature (or unbalanced moment) relations were

determined for the exterior roof level connections of the RC specimen (Fig. 5-20), the

interior roof level connections of the PT specimen (Fig. 5-21), and for the interior floor

level connections of the PT specimen (Fig. 5-22). Calculated yield curvatures for the

slab and column are also indicted on these figures. For negative moment and curvature,

Fig. 5-20(a) (RC-FL2NE-Run4) reveals that slab yielding occurs at close to the

calculated yield curvature, and that the column curvature remains approximately

constant for higher slab curvature values. The results indicate that slab yielding has

occurred without punching failure, as the moment transfer capacity of the slab - column

connection has not been reduced significantly. In contrast, for positive curvatures,

column curvatures begin to drop for higher slab curvatures, indicating that the moment

transfer capacity of the slab - column connection is degrading, which is consistent with

punching failure. Therefore, the results indicate that a punching failure occurred for

FL2NE connection of the RC specimen between t = 12 and 13 sec of Run 4.

The slab curvature versus column curvature plot for the exterior slab - column

connection (Fig. 5-20b; RC-FL2NE-Run4), located opposite to the exterior connection

62
of Fig. 5-20(a) (RC-FL2NW-Run4), displays similar features with punching failure

noted for negative curvatures. For positive moment and curvature, the transferred

moment to the column is constant and relatively small, which is consistent with the

lower moment capacity associated with the slab bottom reinforcement. Column and slab

curvatures for Fig. 5-20 were obtained from displacement gauges.

Figure 5-21 features column and slab curvature data obtained from rebar strain gauges

for an interior connection located at the second story of the PT specimen (PT-FL2NC-

Run4). Significant inelastic slab curvature is observed for both negative bending (Fig.

5-21a) and positive bending (Fig. 5-21b), without a drop in the column curvature,

indicating that punching failure has not occurred (note that the PT specimen was

subjected to more intense motions, i.e., Run 5). In contrast, at the first story interior

connection (PT-FL1NC-Run4), the unbalanced moment transferred from slab to

columns degrades as shown in Fig. 5-22(a) for negative bending and Fig. 5-22(b) for

positive bending. Degradation is characterized by a drop in the unbalanced moment for

increasing slab curvatures during the cycle to peak unbalanced moment, as well as by

noting that the unbalanced moment transfer is reduced for the subsequent cycles to the

same slab curvature for both positive and negative values.

5-6. Drift Capacity of Slab-column Connections at Punching

Relationships for drift capacity at punching versus gravity shear ratio are commonly

derived and compared to the existing database for monotonic and cyclic static tests of

isolated connections. The data are needed to assist in the development of connection

models that account for punching, include the influence of the stud-rails, as well as

address modeling issues associated with stiffness and ductility.

63
Given that columns are typically stiff relative to slabs, the drift ratio is commonly

assumed to be equal to the slab - column connection rotation capacity, with the ultimate

rotation (θu) equal to the sum of the yield rotation (θy) and the plastic rotation (θpl). The

yield rotation for the connections of the test specimens was computed using material

information and equilibrium requirements as discussed in the following paragraphs.

Once the yield rotation was computed, measurements taken during the test were used to

estimate the ultimate rotation, and the plastic rotation was obtained as the difference

between the ultimate and yield rotations.

The yield curvatures (φy) over the flexural transfer width of c + 3h and the column strip

width were computed for slab - column connections of the RC and PT specimens,

respectively, using (5-1) through (5-3).

1
As f s = (c + 3h)( y )( E c ε c ) (5-1)
2

1
As f s + Pps ,cs = (bcs )( y )( Ec ε c ) (5-2)
2

1
As f s + A ps f ps = (bcs )( y )( Ec ε c ) (5-3)
2

Where As is the area of top or bottom steel bars within c + 3h for the RC specimen or

the column strip for the PT specimen, fs is the stress in the bonded steel bars at 0.002

strains (as listed in Table 3-4), c is the dimension of rectangular column parallel to

loading direction, h is the slab thickness, y is the depth to neutral axis, Ec is the secant

concrete modulus of elasticity at 0.45fc’ (Fig. 3-11), εc is the strain of extreme

compression fiber based on linear relation between concrete stress and strain with slope

of Ec, Pps,cs is the average tendon force within the column strip in compression zone, bcs

64
is the width of column strip for the PT specimen, Aps is the area of post-tensioned

tendons in tension zone, and fps is the stress in the unbonded tendons at nominal strength.

For post-tensioned connections, the average tendon force (Pps,cs) was assumed to be

equal to the measured tendon forces at the slab edge, and the concrete compressive

stress was computed assuming that the tendon forces spread out from the slab edge at a

2:1 ratio. The tendon stresses (fps) at punching monitored using load cells were used to

compute the negative yield curvature at interior connections of the PT specimen.

The measured ultimate curvature (φu,dcdt) at punching obtained using displacement

gauges exceeded the yield curvature (φy) at every connection of the RC specimen. To

determine yield rotations for the connections, it is necessary to know the moment

distribution along the slab at first yield of slab reinforcement as well as the location of

the slab inflection point between columns. To determine this information, an

incremental analysis using the effective beam width model (RC, α = 0.77, β = 0.4; PT,

α = 0.63, β = 0.5) for unit acceleration at the base of the model was conducted using

SAP 2000 (Computers and Structures, Inc., Nonlinear Version 7.42). Slab moment

diagrams when yielding of slab reinforcement occurs are shown in Fig. 5-23 and 5-24

for the RC and PT specimens, respectively, for combined earthquake and gravity load

for a slab width of c + 3h (RC) and the column strip (PT). For the RC specimen, a width

of c + 3h was used because slab reinforcement generally does not yield outside this

width, as well as to obtain a lower bound estimate of the yield rotation (and thus, a

lower bound estimate on ultimate rotation capacity). For the PT specimen, banded

tendons act as a one-way beam, such that the compressive resultant spreads out to the

column strip (versus c + 3h); therefore, the width of the column strip is used to compute

65
the yield curvature (φy) and rotation (θy). Since the prestressing forces balance

approximately 70% of dead loads, the moment diagrams for the PT specimen are nearly

linear over the span as shown in Fig. 5-24. The results of the incremental analyses of the

RC (Fig. 5-23) and PT (Fig. 5-24) specimens indicated that slab yielding occurs first in

the bottom reinforcement (positive moment), followed by yielding of top reinforcement

(negative moment), with a shift in the inflection point. Based on the moment diagrams

plotted on Fig. 5-23 and 5-24, yield rotation (θy) values were computed for each

connection. For interior connections, the yield rotation was taken as the average of the

yield rotations on either side of the connection.

The plastic rotation (θpl) for the peak value of unbalanced moment for the RC specimen

was determined using the difference in average slab curvature measured from

displacement gauges anchored to the two threaded rods embedded through the slab

minus the yield curvature (φu,dcdt – φy), multiplied by the distance between the outermost

anchor and the column face (Fig. 5-25) as:

θ pl = φ pl l p = (φ u − φ y )(l p ) ≅ (φ u ,dcdt − φ y )(r2 ) (5-4)

Where r2 is the distance from the outermost anchor point for the displacement

transducer to the column face, lp is the plastic hinge length, φu,dcdt is the average

curvature obtained from the displacement transducer over its gauge length, and φy is the

yield curvature computed using known material relations. For interior connections, the

positive rotation at one side of the column and negative rotation at the other side of the

column were averaged to obtain rotation values.

This process was used because the added information that might have been provided by

66
the rebar strain gauge located one inch from the column face (and between the

displacement gauge anchored approximately 1.5 in. (38.1 mm) and the column face)

was typically not available because the rebar strain gauges had generally failed by this

time. In addition, curvature results from displacement and rebar strain gauges were in

close agreement prior to loss of strain gauge data. It is noted that the length used to

determine average rotations (r2) is typically about 5 in. (50.8 mm), which is within the

range of values expected for the plastic hinge length (lp) of 3.3 in. (83.8 mm) to 6.8 in.

(172.7 mm) using relations suggested by Sawyer (1964), Corley (1967), Mattock (1967).

Values for r1 and r2 are listed in Table 5-1 for all sensors.

For the PT specimen, the curvatures derived from strain gauges located approximately

one inch (50.8 mm) from the column face were greater than the average curvatures

derived from displacement gauges; therefore, both measurements were used in

calculation of the plastic rotation (θpl). For this case, the plastic rotation was calculated

as (Fig. 5-26):

θ pl = φ pl l p = (φ u − φ y )(l p ) ≅ (φ u , sg − φ y )(r1 ) + (φ u ,dcdt − φ y )(r2 − r1 ) (5-5)

Where r1 is the distance from the inner threaded rod used to anchor the displacement

gauges and the column face, and φu,sg is the average curvature obtained using the rebar

strain gauge data.

The ultimate rotation capacity (θu), which is approximately equal to the drift ratio at

punching, is the sum of the yield rotation (θy) and plastic rotation at the peak (θpl). Table

5-1 and 5-2 summarize the calculated rotations and the resulting drift capacity of each

connection. Gravity shear ratios given in Table 5-2 and 5-3 were computed using the

67
concrete strength obtained from the core sample tests. Prior test results, as summarized

by Pan and Moehle (1989) and Robertson et al. (2002) are compared with the data

contained in Table 5-2 and 5-3, and Fig. 5-27.

Drift capacities for the interior connections range between 2.14 % and 2.37 % (average

value = 2.26 %) for the RC specimen and 3.07 % and 3.41 % (average value = 3.24 %)

for the PT specimen. The drift capacities for the exterior connections are larger than the

values for the interior connections, particularly for the PT specimen. Drift values for the

PT specimen are larger, in part, because the system is more flexible (e.g., see Table 5-3).

The plots indicate that the drift ratios at punching failure obtained from the shake table

tests are lower than observed for isolated connections with stud-rails tested under quasi-

static load or displacement histories. The results agree fairly closely with trends

observed for isolated connection tests without shear reinforcement.

Results from strain gauges mounted on stud-rails indicate that the stud-rails did not

reach yield. For this reason, the ductility of the slab - column connections of the RC and

PT specimen is less than observed in prior tests. The tests of isolated connections with

stud-rails conducted by Elgabry and Ghali (1987), and Megally and Ghali (2000) under

static monotonic or cyclic lateral load achieved significantly larger drift ratios than the

tests described herein. The higher drift ratios may be influenced by yielding of stud-rails

prior to observed punching failures.

5-7. Monitoring and Variation of Post-tensioning Forces

Tendon forces obtained during the post-tensioning process from load cells (LC)

mounted at tendon ends are given in Table 5-4. The first post-tensioning and third post-

68
tensioning were conducted to support slab self weight and slab self weight and the total

weight of the lead, respectively (see Chapter 4, Section 1). Results from LC # 2 and LC

# 4 (Table 5-5) installed to monitor the compressive force at the tendon dead end and

the jacking end, respectively, indicate that the rate of friction loss over the tendon is

0.00534 kips/in. (9.35 N/cm) at the completion of the first post-tensioning (Fig. 5-28).

Once the tendons were stretched during the first post-tensioning, the rate of friction loss

over the tendon length decreased to 0.00345 kips/in. (6.04 N/cm) at the completion of

the third post-tensioning. Anchor set loss (∆L) of 0.0105 in. (0.27 mm) and friction loss

coefficient (kx+µα) of 0.073 were obtained from the readings during the third post-

tensioning using (5-6) and (5-7). Based on these findings, the resulting distribution of

post-tensioning force along the two span slabs prior to and after anchor set is assumed

to be linear as shown in Fig. 5-29. Since the dead ends and jacking ends were alternated,

the average post-tensioning force per tendon at both interior and exterior connections

can be estimated as approximately 9.9 kips/tendon (44 kN/tendon) prior to the dynamic

tests.

E ⋅ ∆L ⋅ d
∆f = (5-6)
36 L

Where ∆f is the change in stress due to anchor set, d is the friction loss in length L; L is

the length to point where loss is known, E is the modulus of elasticity, and ∆L is the

anchor set loss.

The post-tensioning force at distance x along the strand is computed as:

Px = Ps /(1 + kx + µα ) (5-7)

Where Ps is the post-tensioning tendon force at the jacking end, Px is the post-tensioning

69
tendon force at distance x from the jacking end, k is the wobble friction coefficient per

meter of post-tensioning tendon, µ is the curvature friction coefficient, α is total angular

change of post-tensioning tendon profile in radians from tendon jacking end to location

x.

During Runs 4 and 5 for the PT specimen, tendon stress increased by a maximum of

22.9 ksi (158 MPa) to 30.6 ksi (211 MPa), respectively, based on data obtained from the

donut-shaped load cells. These values represent the difference between the initial post-

tensioning force prior to the test and the peak stress of tendons during each test.

According to the ACI 318 Building Code (“Building”, 2002; see Section 18.7.2), the

increase in the effective tendon stress for the PT specimen is determined as:


fc
f ps = f se + 70 + (5-8)
300 ρ p

Where fps is the stress in the unbonded tendons at nominal strength, and shall not taken

greater than fpy nor greater than (fse + 200); fpy is the yield stress in the unbonded

tendons, fse is the effective prestress in prestressed reinforcement after allowance for all

prestress losses, MPa; and ρp is the ratio of prestressed reinforcement, dp is the distance

from extreme compression fiber to centroid of prestressed reinforcement. The resulting

increase in the effective tendon stress using (5-8) is approximately 19.6 ksi (135 MPa),

which is lower than the peak stresses measured during the test. Test results appear to

indicate that the ACI 318 equation is conservative.

Tendon stresses at the ultimate load carrying capacity of post-tensioned slabs have been

reported by numerous researchers (Trongtham and Hawkins, 1977; Burns and

70
Hemakom, 1977; 1985; Kosut et al., 1985; Dilger and Shatila, 1989; Foutch et al., 1990;

Qaisrani, 1993). In general, the peak stresses reported in these studies are less than those

predicted by (5-8), whereas the measured peak tendon stresses using load cells in this

test are higher than the values computed using (5-8).

Equation (5-8) was developed primarily from beam tests. Naaman and Alkhairi (1992)

conducted a review of tendon stresses at ultimate for 143 beams where flexural failures

were reported. The review revealed that the tendon stress at ultimate predicted by ACI

318-02 (“Building”, 2002) is conservative. In contrast to beam test results, results from

tests of slab - column connections noted in the prior paragraph have shown that (5-8)

over estimates the tendon stress at punching failure, whether the slab reinforcement

yields, or does not yield. The lower tendon stresses reported for slabs failing due to

punching relative to beams failing in flexure is reasonable. As well, in a majority of

these tests, gravity shear stresses at the connection were relatively high compared with

the ratios used for the PT specimen tested in this study. Given the lower gravity shear

ratio, the drift capacity of the PT specimen at punching failure was relatively large

compared to the drift capacity of previous test specimens, resulting in higher tendon

stresses at ultimate than noted in prior tests.

71
72
6. Analytical Studies
6-1. Introduction

As discussed in Chapter 2, two general approaches exist for modeling and analysis of

slab - column frames, the effective beam width model and equivalent frame model. The

initial work (Peabody, 1948) on the equivalent frame analysis approach was done for

gravity loads only; however, the approach was later extended to cover combined gravity

and lateral loading (Vanderbilt and Corley, 1983). In this chapter, analytical studies are

conducted using appropriate models to compare the responses obtained with analytical

models with responses obtained from the shake table tests.

6-2. Description of Analytical Models

In the following sections, the models developed for use in comparing analytical and

experimental results are described in detail. In specific, details are provided on the

assignment of stiffness, strength, and ductility capacity, including punching failure of

slab-column connections.

The OpenSees (http://opensees.berkeley.edu/OpenSees) platform was used to conduct

the analytical studies. The structure was modeled as a plane frame as shown in Fig. 6-1.

The column and the slab were modeled using beam or beam-column elements. Rigid

end zones were used and assumed to extend from the center of the column to the face of

the column (101.6 mm; 4 in.). Test results indicated that the base of the footings

supported on the load cells experienced significant rotation that was strongly correlated

with moment (see Fig. 5-2); therefore, rotational springs were included in the model at

73
the base of rigid column footings to allow direct comparison between experimental and

analytical results. A spring constant of 50,000 k-in/rad (5,650 kN·m/rad) was

determined from the experimental results (Fig. 5-2).

Slab and slab-column connection properties:

Stiffness properties for the slab beams used in the analytical models were determined

using the effective slab width model of Pecknold (1975). The slab - column connections

were assumed to be rigid, as studies indicate that use of this assumption tends to

produce better results (Allen and Darvall, 1975). Given the column and slab dimensions,

α-values of 0.8 and 0.65 were selected for the RC and PT specimens, respectively. As

noted later, some variation of the α-values was considered in the analyses to address the

sensitivity of the results to the α-values selected.

The influence of cracking on the slab stiffness was addressed by applying a stiffness

reduction factor β (e.g., see Moehle and Diebold, 1984). In general, the post-yield

flexural stiffness of the slab was assumed to be in the range of 0 to 20% of the fully-

cracked elastic stiffness (My/φy).

The yield and nominal moment strengths of the effective slab widths were calculated

using well-known procedures, i.e., an assumed linear compressive stress distribution at

yield moment and a Whitney Stress Block at ultimate moment. Moment capacities for

various slab widths were computed, including: (1) the width of the column strip, since

100% of the earthquake moment was assigned to this width, and (2) the transfer width

of c + 3h, since flexural reinforcement for a fraction of the column strip moment is

required to be placed within this width. Post-yield strength is set to an arbitrary value,

74
based on the assumed post-yield stiffness. Initially, the slab beams were assumed to

have infinite ductility.

Gravity loads tributary to the column strip were included in the model. Additional point

loads were applied directly to the columns to account for the tributary gravity load from

the middle strips.

Potential yielding or failure mechanisms within the slab and at the slab - column

connection were considered. Three specific mechanisms were modeled as depicted on

(Fig. 6-2):

(1) The potential for punching failure prior to yielding of slab flexural reinforcement

was modeled using rigid plastic springs as shown in Fig. 6-2(b). Punching failure results

when the shear stress due to combined gravity shear and eccentric shear due to

unbalanced moment reaches the shear capacity defined by ACI 318-02 at any point on

the critical section. Punching failure is modeled as a sudden drop in the connection

moment capacity to a residual value (e.g., zero). Due to the presence of continuous

bottom reinforcement through the column, a shear transfer mechanism was maintained

at the slab - column connection to transfer gravity shear to the column after punching

failure.

(2) The specimens were designed such that flexural yielding within the slab transfer

width of c + 3h was expected prior to punching failure; however, ultimately, punching

failure was expected. Once flexural yielding occurs within c + 3h, additional

unbalanced moment capacity is associated with the residual capacity in eccentric shear;

therefore, after yielding, γv is set to 1.0 and additional unbalanced moment transfer is

75
allowed until the residual capacity is exhausted. The slope of the post-yield relation was

arbitrarily set by defining the rotation associated with the unbalanced moment capacity

(flexure and eccentric shear) to the experimental derived best-fit relation for drift

capacity at punching (Chapter 5, Section 6). To account for punching failure, the

moment-rotation relation for the slab - column connection was modified as shown on

Fig. 6-2(b), such that the connection moment capacity drops to a residual value once the

punching rotation is reached.

(3) Flexural yielding of the slab reinforcement within the column strip was modeled.

Moment values were computed using material information, assuming plane sections

remained plane after loading, and imposing equilibrium. Post-yield relations were

computed using a “plastic-hinge” model with 2% strain hardening, as depicted in Fig. 6-

2(c). Punching failure was eventually expected when the slab rotation exceeds the

rotation capacity from experimental results derived in Chapter 5, Section 6.

Slab properties used for the models of the test specimens are summarized in Table 6-3.

Column properties:

Column stiffness and strength were evaluated two ways: (1) by direct use of the

measured material relations using a fiber model, and (2) by computing moment versus

curvature relations and assigning a bilinear moment rotation spring at the column ends.

Material stress-strain relations used were based on test data and are shown in Fig. 6-6.

The modified Kent and Park model (Park et al., 1982) was used to represent the stress –

strain behavior of the cover and core concrete. This model is relatively simple, but

includes sufficient detail to address the influence of stiffness variation at moderate

76
concrete compressive stress levels on computed responses. A peak unconfined concrete

compressive stress of 23.1 MPa (3.355 ksi), with a strain at peak stress of 0.0024, were

used to match the test results for the concrete. The dependence of column stiffness and

moment capacity on column axial load is considered directly using (1). For (2), this

dependence was considered by computing moment versus curvature relations for gravity

load and gravity load +/– the maximum earthquake shears that could develop at the slab

- column connection. Relations and bilinear fits to the relations are plotted on Fig. 6-3

for interior and exterior columns at the first and second stories. The effective cracked

stiffness value obtained from these relations varied between 0.38EIg and 0.42EIg for the

exterior columns and 0.45EIg for the interior columns. To simplify the model, values of

0.4EIg was used for the interior and exterior columns. Column properties used for the

models of the test specimens are summarized in Table 6-1.

6-3. Analytical Models: Low-to-moderate Level Responses

Initial efforts focused on the use of linear elastic, on nearly linear elastic models, to

compare with low-level and moderate-level responses recorded during Runs 1 and 2 for

the RC and PT specimens. Damping ratios used for the analyses, 3.05% (RC) and

2.65% (PT), were determined from free vibration tests conducted prior to Run 2-2 (RC)

and Run 2 (PT), respectively. Responses for six analytical models were evaluated:

(1) Model A: Uncracked column and uncracked effective slab width (Icol and αIslab)

(2) Model B: Uncracked column and cracked effective slab width (Icol and αβIslab)

(3) Model C: Cracked column and cracked effective slab width (0.4Icol and αβIslab)

(4) Model D: FEMA 273 recommended stiffness values (0.7Icol and 0.35KfpIslab where

Kfp=1.0, 0.8 for interior and exterior joint, respectively)

77
(5) Model E: Fiber column model and uncracked effective slab width (αIslab)

(6) Model F: Fiber column model and cracked effective slab width (αβIslab)

Where Icol and Islab represent the gross section inertia for the column and slab effective

width, respectively. The effective width coefficient α was taken as 0.8 for the RC

specimen and 0.65 for PT specimen based on results reported by Pecknold. (1975).

Coefficient β is a stiffness reduction factor to account for cracking. Vanderbilt and

Corley (1983) suggested using a β-value of 1/3 for an equivalent frame model, whereas

Moehle and Diebold (1984) suggest using values between 1/3 and 1/2 (see Section 2-1).

In this report, β-values of 1/3 and 1/2 were used for the RC and PT specimens,

respectively. The higher value for the PT specimen reflects that reduced level of

cracking due to the presence of the post-tensioning reinforcement.

Models A - D are linear elastic, whereas Models E and F include minor non-linearity,

because the column was modeled using fiber elements. Using the fiber model, the

column stiffness is derived directly from the material stress – strain relations; therefore,

change in concrete flexural stiffness at low-to-moderate stress levels is modeled directly.

Maximum concrete compressive strains reached for Run 2-2 (RC) or Run 2 (PT) were

less than the strain at peak stress; therefore, Run 2-2 (RC) or Run 2 (PT) comparisons

for Models E and F consider only the nonlinearity associated with responses that do not

exceed the peak compressive stress for concrete in compression.

6-3-1. Period Comparisons

Fundamental periods obtained from the various models and period ratios equal to the

model fundamental period divided by the test structure fundamental period are

78
summarized in Table 6-1. The fundamental period of the test structure was obtained

from two ways: (1) from the free vibration tests conducted prior to Run 2-2 and Run 2

for the RC and PT specimen, respectively, and (2) from Fourier amplitude spectra of the

roof acceleration response histories during Run 2-2 (RC) and Run 2 (PT) (moderate

level shaking expected to produce concrete cracking but no reinforcement yielding).

RC Specimen: Results for Model A, without slab or column cracking, indicate a period

ratio is 1.32; however, if slab cracking is included (Model B), then the period ratio is

1.04, indicating that the stiffness is overestimated by approximately (1.04)2 = 1.08. The

period ratio for Model C, including both slab and column cracking is 0.98. Results for

Models B and C are consistent with expectations, and indicate that these models

reasonably represent the global structural response. The period ratio for Model D is 1.06,

indicating that the model is a little stiff. For the models using fiber element of column,

Model E shows a period ratio of 1.30 indicating that the model is too stiff. A period

ratio of 1.04 was computed for Model F, indicating good correlation between the

analytical and experimental results for the analyses utilizing the column fiber element

model. The result for Model F suggests that the Model C overestimates the influence of

cracking on column stiffness reduction.

PT Specimen: For the PT specimen, including slab cracking with a β-value of 1/2

(versus 1/3 used for the RC specimen) results in good agreement between measured and

computed fundamental periods, with period ratios of 1.0 and 0.94 for Models B and C,

respectively. The period ratio for Model D of 0.97 indicates that this model also

represents the stiffness of the test structure reasonably well. A period ratio of 1.00 was

79
computed for Model F, which again suggests that Model C overstates the impact of

cracking on column stiffness reduction.

6-3-2. Base Shear and Top Relative Displacement Comparisons

Results for base shear versus top relative displacement for the RC and PT specimens are

presented in Fig. 6-4. For the RC specimen, test results for Run 2-2 compare favorably

with analytical results for Model C (0.4Icol and αβIslab = 0.27Islab). For the PT specimen,

Model C (0.4Icol and αβIslab = 0.33Islab) and D (0.7Icol and 0.35KfpIclab) closely match the

test results.

Top relative displacement histories for Models B and C are compared in Fig. 6-5 and

Fig. 6-6 for the RC specimen subjected to Run 2-2. In Fig. 6-6(a), results compare

reasonably well, although the analytical model is too stiff. Results presented in Fig. 6-

6(b) match reasonably well, particularly from 15 to 17 seconds. Results for the PT

specimen (Fig. 6-5 and Fig. 6-7) subjected to Run 2 compare very favorably. In general,

the experimental results have lower peak responses, indicating that the damping ratio

used for the analytical models may be too low. This is especially true for the response of

the PT model. The better results for the PT specimen may be influenced by the fact that

rotation at the base of the footings was measured directly for the PT specimen, whereas

for the RC specimen, results for the PT specimen were used to estimate footing rotation

for the RC specimen.

Use of a fiber model for the columns improved response comparisons between

experimental and analytical relations as shown in Fig. 6-8 (relative to the results

presented in Fig. 6-4).

80
Top relative displacement histories for Model F are compared in Fig. 6-9 for the RC and

PT specimens subjected to Run 2-2 and Run 2, respectively. In Fig. 6-9(a) for the RC

specimen, period and peak values compare reasonably well, although the overall

comparison is marginal. Results for the PT specimen in Fig. 6-9(b) compare very

favorably, although the results do not appear to be any better then the results presented

for the Model C.

6-4. Analytical Models: Nonlinear (Yielding) Responses

In the preceding sections, comparisons were presented for results obtained from the RC

and PT specimens and analytical models for low-to-moderate levels of shaking where

essentially elastic behavior was anticipated. For greater intensity shaking, the nonlinear

behavior of the specimens must be modeled to allow comparisons between experimental

and analytical results. As noted in Section 6-2, nonlinear behavior due to yielding of

slab reinforcement within the column strip or the transfer width of c + 3h adjacent to the

slab - column connections is considered. Drift capacities at which punching failures of

the slab - column connection are expected to occur are based on the experimental results

(Chapter 5, Section 6). Failure mechanisms of slab - column connections or slabs across

the column strip are categorized into the following three potential items.

(1) Punching failure prior to yielding of slab flexural reinforcement (item 1, Fig. 6-2b).

(2) Punching failure at the ultimate rotation of the connection spring (θu) after flexural

yielding within the slab transfer width of c + 3h (item 2, Fig. 6-2b).

(3) Flexural yielding of the slab reinforcement within the column strip, followed by

punching failure at the critical slab rotation (item 3, Fig. 6-2c).

81
Additional details on the nonlinear modeling of the slab elements, slab - column

connection elements are discussed in the following paragraphs. For the column

elements, the fiber model was used.

6-4-1. Punching Failure prior to Slab Flexural Yielding

Vg
Punching failure occurs when the sum of direct gravity shear stress ( ) and shear
bo d
stress induced by a fraction of unbalanced moment transferred by eccentric shear
γ v M ecc,unb c
( ) reaches to shear stress capacity (vn), where c is the distance from the
Jc
centroid to the critical section that results in the smallest value of Mecc,unb and vn is the

nominal shear capacity at the connection (see Chapter 2, Section 2-2). The nominal

shear strength vn within the shear reinforced region is vn = vc + vs and the nominal shear
2 ′
strength outside the shear reinforced zone is vn = vc. The upper limit (vn = f c MPa )
3
of nominal shear strength vn within the shear reinforced region suggested in ACI
1 ′
421.1R-99, instead of f c MPa , is used for the PT specimen.
2

For this case, the fraction of unbalanced moment transferred by flexure (γfMecc,unb) does

not exceed the yield moment of slab reinforcement within c + 3h for the exterior

connection or the sum of the positive moment on one side and negative moment on the

other side for the interior connection. If the yield capacity of the connection spring is

determined using the eccentric shear stress model (i.e. no flexural yielding within c +

3h), the connection spring drops without ductility as shown in Fig. 6-2(b) (item 1).

For most connections of the RC and PT specimen, punching failure is not expected prior

to flexural yielding within c + 3h. The positive moment transfer capacity for the exterior

82
connections of the PT specimen is limited by punching failure just prior to yielding

within c + 3h and the column strip (which are almost the same in this case, since 75 %

of all the reinforcement is within c + 3h). In the initial analyses presented in this chapter,

punching failure prior to connection yielding was suppressed; therefore, punching

failure was governed by inelastic slab rotations reaching the critical rotation θu based on

the experimental results. Additional studies are planned to examine the impact of this

assumption, as well as variation of other parameters such as material strengths (which

impact punching strength).

6-4-2. Punching Failure after Slab Flexural Yielding within c + 3h

The capacity to transfer unbalanced moment from the slab to the column was modeled

using a rigid plastic (connection) spring as shown in Fig. 6-2. The yield moments of the

slabs framing into a connection over a transfer width of c + 3h were computed for both

interior and exterior connections. For an interior connection, the yield moment of the

spring is determined by summing the positive moment on one side of the connection

and negative moment on the other side of the connection.

+ −
M y ,unb , f = M y ,c +3h and M y ,c +3h : for exterior connections (6-1)

±
M y ,unb , f = M y ,c + 3h + M y ,c + 3h : for interior connections
m
(6-2)

where My,c+3h+ and My,c+3h– are computed using the actual material properties in

accordance with Chapter 5, Section 6. Since only a fraction of unbalanced moment is

assigned to flexure (e.g., γfMunb), the yield capacity for unbalanced moment transfer in

the connection spring (My,unb) is determined from the following relation:

83
M y ,unb , f = γ f M y ,unb (6-3)

The remaining fraction of unbalanced moment (γvMy,unb) assigned to eccentric shear is

less than that required to result in punching failure using the eccentric shear stress

model. Yield capacities for transfer of unbalanced moment at slab - column connections

and within the column strip are summarized in Table 6-2. In most cases, the yield

moment of the connection spring does not exceed the yield moment of the slab (exterior

connection) or slabs (interior connection) within the column strip framing into the

connection; therefore, yielding of the connection spring is expected prior to flexural

yielding of the slabs across the entire column strip. In (6-3), the fraction of unbalanced

moment assigned to flexure (γf) is computed based on ACI 318-02, Section 13.5.3.3

(“Building”, 2002).

The ultimate capacity of the connection spring to transfer unbalanced moment is

determined by determining the unbalanced moment that exhausts the nominal moment

strength of slab (exterior) or slabs (interior) within c + 3h (because yielding occurs prior

to punching); therefore, M n ,unb, f + M n,unb,v = γ f M n,unb + γ v M n,unb = M n,unb and the

residual unbalanced moment transfer capacity associated with eccentric shear only

(∆Munb,v). After yielding, γv is assigned a value of unity and the transfer moment

increases until: (1) the punching shear limit is reached, or (2) the column strip yields

and the slab rotation reaches a value that results in punching failure (described in

Section 6-4-3). The nominal strength for unbalanced moment transfer (Mn,unb) is

calculated using the same procedure as (6-1), except using the Whitney Stress Block

method with the experimentally determined yield stress (fy) of reinforcement and the

mean compressive strength of the concrete core samples (fc’) (as listed in Table 3-1 and

84
3-4). The average of core sample tests for each specimen is used to calculate the

incremental unbalanced moment (∆Munb,v) that causes the eccentric shear capacity of the

connection to be reached.

The post-yield stiffness depends on the ultimate capacity of unbalanced moment

transfer at punching (Mn,unb,f + Mn,unb,v + ∆Munb,v = Mu,unb) and the connection rotation

capacity from the experimental results (θu) as shown in Fig. 6-2(b) (item 2). For the PT

specimen, since tendons are banded and most bonded reinforcement is within c + 3h,

slab flexural yielding is modeled only by using column strip springs, such that failure

modes of the PT specimen are determined as either item 1 or item 3.

6-4-3. Punching Failure after Slab Flexural Yielding within Column Strip

Flexural yielding in the slab adjacent to the slab - column connection is considered on

either side of the connection (item 3, Fig. 6-2c). Punching failure is modeled by

assuming the moment capacity of the slab drops to a residual (e.g., zero) capacity once a

critical rotation is reached. Since a plastic rotation of an exterior or interior connection

consists of a plastic rotation of the connection spring and a plastic rotation of the

column strip spring (exterior) or the average plastic rotation of the column strip springs

on each side (interior), the critical rotation is determined as the sum of the plastic

rotation within the connection spring prior to yielding of the column strip spring and the

plastic rotation of the column strip spring. Therefore, the residual plastic rotation

capacity of the column strip spring (where connection spring yielding occurs first) is

calculated by subtracting connection spring rotation within c + 3h from the critical

punching rotation obtained from experimental best-fit results. Yield and punching

rotation values obtained are summarized in Table 5-2 and 5-3 for each specimen.

85
Yielding capacity of the column strip spring (Mfy,cs) is modeled separately from the

ultimate capacity of the connection spring (Mn,unb,f + Mn,unb,v + ∆Munb,v), such that

punching failure of the column strip spring can occur prior to punching failure of the

connection spring (or vice versa).

6-4-4. Response of Static Push-over Analyses

At this time, nonlinear analysis results are presented only for static push-over analyses,

as shown Fig. 6-10, for the RC and PT specimens, respectively. The push-over curves

plotted in Fig. 6-10 are the analytical results conducted with 0.5:1 and 2:1 for the

relative magnitudes of the lateral loads applied at the second and first floor levels,

respectively. The lateral load distribution ratios of 2:1 and 0.5:1 resulted in better

predictions in linear and nonlinear range, respectively, which are consistent with the

discussion in Chapter 2, Section 6 (Fig. 2-15) for prior results presented by Moehle and

Diebold (1984).

Test results are plotted for Runs 2-2 and 4 (RC) or Runs 2, 4, and 5 (PT), to allow

comparison for both low-to-moderate response levels, as well as for significant

nonlinear responses. The pushover relations for the RC (Fig. 6-10a) and PT (Fig. 6-10b)

specimens compare quite well with the response envelop for the test results. The one

discrepancy for both cases occurs during the first large excursion into the nonlinear

range for positive loading. For this excursion, loading rate may have resulted in

additional capacity that is not considered in the model. Future studies may be conducted

to address this issue.

86
7. Summary and Conclusions
7-1. Summary

An investigation into the lateral load response of reinforced concrete flat plate frames

utilizing stud rails for shear reinforcement at the slab - column connections was carried

out. One of the specimens consisted of a conventional reinforced concrete flat plate (RC

specimen), whereas the other specimen consisted of nominally reinforced flat plate with

post-tensioning reinforcement (PT specimen). The two-bay by two-bay, two-story

specimens were approximately one-third scale representations of typical slab - column

frames constructed in moderate-to-high seismic zones in the United States and Japan.

The specimens were subjected to gravity loads and increasing intensity of uniaxial base

acceleration histories on the shake table at the Earthquake Engineering Research Center

at UC Berkeley’s Richmond Field Station.

Details of the research program and the research results are described in six chapters.

Chapter 2 provides a literature review, followed by Chapter 3, which presents details on

the design and construction of the specimens. Details of the instrumentation and shake

table motions are described in Chapter 4. Chapters 5 and 6 present preliminary results

from the investigation. Detailed evaluations of the experimental results, including an

assessment of the consistency of the data collected, are included in Chapter 5, whereas

Chapter 6 presents comparisons between results obtained with analytical models with

the results obtained from the tests of the two specimens.

7-2. Conclusions

87
Based on the information presented in Chapters 1 through 6, a number of observations

and conclusions are noted. General findings include:

(1) Post-tensioned slab - column frames incorporating shear reinforcement within the

slab - column connection region are commonly used for so-called non-participating

systems or gravity systems in structures constructed in high seismic zones. The use of a

post-tensioned slab allows for higher slab span-to-depth ratios, whereas the use of shear

reinforcement increases the shear transfer capacity in the slab - column connection

region thereby eliminating the need for a column capitol or drop panel.

(2) Although the use of non-participating slab - column frames is common, relatively

little experimental data are available to assess the dynamic responses of such systems,

particularly for post-tensioned systems.

(3) Deterioration of the moment capacity at the slab - column connections occurred

during the tests; however, lateral drift ratios of 3% and 4% were achieved for the RC

and PT specimens, respectively, with relatively little loss of lateral load capacity. The

results indicate that the slab - column frames can be designed to have sufficient drift

capacity to be used as a non-participating system in conjunction with shear walls or an

alternative lateral force resisting system in high seismic regions, or as a primary lateral

force resisting system in low-to-moderate seismic regions.

More specific observations and conclusions, based on evaluation of the experimental

data or comparisons between experimental results and the results from analytical

models include:

(4) Evaluation of column moment values obtained using three different approaches was

88
in reasonably close agreement, providing confidence that the measured responses were

reliable. The data assessment reveals that column moments derived from the load cells

under the column footings and the rebar strain gauges at the base of the columns agree

closely. A review of data used to derive column curvatures at the base of the first story

column from rebar strain gauges and displacement gauges, as well as rebar strain gauges

placed symmetrically on slab reinforcement, indicated that the data collected during the

tests were consistent and of good quality.

(5) Based on measured rebar strains, yielding of slab reinforcement was limited to the

effective width of c2 + 3h or c2 + 2ct as defined in ACI 318-02, and observed cracks at

exterior connections were consistent with the 45 degree yield lines assumed in ACI 318-

02. For the PT specimen, flexural cracks were observed across the full slab width at the

end of testing, indicating that the post-tensioning was effective in developing the full

nominal moment strength of the slab.

(6) Based on the observed damage at slab - column connections, it was concluded that

connection punching had occurred during the test program, despite the somewhat

limited connection damage observed compared with tests conducted under slowly

varying loads on specimens of similar scale. The limited damage at punching failure is

in sharp contrast to that observed for tests of reinforced concrete flat plate tests without

shear reinforcement (e.g., compare Fig. 1-4 and Fig. 5-16).

(7) Punching at slab - column connections was evaluated by plotting slab curvature

versus column curvature (or unbalanced moment) relations. Punching failures were

noted in cases where column curvature (or unbalanced moment) diminished as slab

curvatures increased. Identification of slab and column curvatures (and moments) at

89
which punching occurred allowed the determination of connection drift capacities at

punching failures.

(8) Connection punching failures for interior connections were determined to occur for

drift levels between 2.13% and 2.36%, with an average value of 2.25 % for the RC

specimen, and 3.07% and 3.41% with an average value of 3.24% for the PT specimen.

Drift levels at punching failure for the exterior connections were substantially larger

than those for interior connections, especially for the post-tensioned connections, where

the drift levels for exterior connections were approximately 1.5 times those for the

interior connections.

(9) Connection drift levels associated with punching failure for the two test specimens

are substantially less than values obtained from tests of isolated slab - column

connections. The results agree fairly closely with trends observed for isolated

connection tests without shear reinforcement. Since the stud rails did not reach yield,

the ductility of the slab - column connections of the RC and PT specimens are less than

observed in prior static tests which revealed yielding of stud rails prior to observed

punching failures.

(10) Based on results obtained from load cells, peak tendon stresses of 22.34 ksi (154

MPa) to 30.93 ksi (213 MPa) were measured in the unbonded tendons during the

dynamic tests. The peak values exceed those measured in tests of connections with

relatively high gravity shear ratio subjected to quasi-static loading, suggesting that the

tendon stress at nominal strength used in ACI 318-02 (“Building”, 2002; see Section

18.7.2) may be conservative.

90
(11) Analytical models including column cracking with an effective width factor α of

0.8 and 0.65, and a cracking factor β of 1/3 and 1/2, for the RC and PT specimens,

respectively, resulted in good correspondence between experimental obtained and

analytical computed fundamental periods and top level displacement responses for low-

to-moderate levels of shaking (i.e., prior to yield). Analytical results for the PT

specimen showed better agreement with experimental results then did the RC specimen,

possibly due to the influence of footing rotation which was directly measured for the PT

specimen.

(12) Analytical models were created using nonlinear springs to model slab yielding and

connection punching failures to compare with measured top level displacement versus

base shear relations. Experimental and analytical results were compared for Run 4 for

both specimens. In general, the analytical models reasonably predicted the experimental

results. The one discrepancy was observed during a single, substantial excursion into

the nonlinear range, where rate effects may have played a role. Additional work is

needed to produce more detailed comparisons between experimental and analytical

results.

91
92
References

ACI-ASCE Committee 352. (1988). Recommendation for Design of Slab-Column


Connections in Monolithic Reinforced Concrete Structures (ACI 352.1R-89). ACI
Structural Journal, V. 85, No. 6, pp 675-696.
ACI-ASCE Committee 421. (1999). Shear Reinforcement for Slabs (ACI 421.1R-99).
American Concrete Institute, Detroit, MI.
ACI-ASCE Committee 423. (1996). Recommendation for Concrete Members Prestressed
with Unbonded Tendons (ACI 423.3R-96). American Concrete Institute, Detroit, MI.
Allen F. H., and P. LeP. Darvall. (1975). Lateral Load Equivalent Frame. ACI Structural
Journal, Proceedings V. 74, No. 7, pp 294-299.
Ascheim M., and J. P. Moehle. (1995). Four-Story Flat-Slab Building in Sherman Oaks.
Preliminary Building Report to ATC-33 Case Study.
Banchik C. A. (1987). Effective Beam Width Coefficients for Equivalent Frame Analysis
of Flat Plate Structure. ME thesis, Department of Civil and Environmental Engineering,
University of California at Berkeley.
Building Code Requirements for Structural Concrete. (ACI 318-77). American Concrete
Institute, Detroit, MI.
Building Code Requirements for Structural Concrete. (ACI 318-83). American Concrete
Institute, Detroit, MI.
Building Code Requirements for Structural Concrete. (ACI 318-99). American Concrete
Institute, Detroit, MI.
Building Code Requirements for Structural Concrete. (ACI 318-02). American Concrete
Institute, Detroit, MI.
Burns N. H., and R. Hemakom. (1977). Test of Scale Model Post-Tensioned Flat Plate.
Journal of the Structural Division, ASCE, V. 103, No. ST6, pp 1237-1255.
Burns N. H., and R. Hemakom. (1985). Test of Post-Tensioned Flat Plate with Banded
Tendons. Journal of the Structural Division, ASCE, V. 111, No. 9, pp 1899-1915.
Corley W. G. (1967). Rotational Capacity of Reinforced Concrete Beams. Journal of
Structural Division, ASCE, V. 92, No. ST5, pp 121-146.
Computers and Structures, Inc. (2000). SAP 2000. Nonlinear Version 7.42, Berkeley,
California.
Decon, Inc. (1996). Decon® Studrail® Design Manual. Ontario, Canada.

93
Dilger W. H., and M. Shatila. (1989). Shear Strength of Prestressed Concrete Edge Slab-
Column Connections with and without Shear Stud Reinforcement. Canadian Journal of
Civil Engineering, V.16, pp 807-819.
Di Stasio J., and M. R. Van Buren. (1960). Transfer of Bending Moment Between Flat
Plate Floor and Column. ACI Structural Journal, V. 57, No. 3, pp 299-314.
Elgabry A. A., and A. Ghali. (1987). Tests on Concrete Slab-column Connections with
Stud-Shear Reinforcement Subjected to Shear-Moment Transfer. ACI Structural Journal,
V. 84, No. 5, pp. 433-442.
Foutch D. A., Gamble W. L., and H. Sunidja. (1990). Tests of Post-Tensioned Concrete
Slab-Edge Column Connections. ACI Structural Journal, V. 87, No. 18, pp 167-179.
Hayes J. R., Foutch D. A., S. L. Wood. (1999). Influence of Viscoelastic Dampers on the
Seismic Response of a Lightly Reinforced Concrete Flat Slab Structure. Earthquake
Spectra, V. 15, No. 4, pp 681- 707.
Hammill N., and A. Ghali. (1994). Punching Shear Resistance of Corner Slab-Column
Connections. ACI Structural Journal, V. 91, No. 6, pp. 697-707.
Hanson N. W., and J. M. Hanson. (1968). Shear and Moment Transfer between Concrete
Slabs and Columns. Journal, PCA Research and Development Laboratories, V. 10, No. 1,
pp. 2-16.
Harris H. G., White R. N., and G. M. Sabnis. (1999). Structural Modeling and
Experimental Techniques 2ed. CRC Press.
Hawkins N. M. (1971). Shear and Moment Transfer Between Concrete Flat Plates and
Columns. Progress Report on National Science Foundation Grant No. GK-16375,
Department of Civil Engineering, University of Washington at Seattle.
Hawkins N. M., and W. G. Corley. (1971). Transfer of Unbalanced Moment and Shear
from Flat Plate to Columns. Paper SP30-7, Cracking, Deflection and Ultimate Load of
Concrete Slabs Systems, pp. 147-176.
Hawkins N. M., Mitchell D., and M. S. Sheu. (1975). Reversed Cyclic Loading Behavior
of Reinforced Concrete Slab Column Connections. Proceedings, U.S. National
Conference on Earthquake Engineering, Ann Arbor, MI, pp. 306-315.
Hawkins N. M., Mitchell D., S. N. Hanna. (1975). The Effects of Shear Reinforcement on
the Reversed Cyclic Loading Behavior of Flat Plate Structures. Canadian Journal of Civil
Engineering, V.2, pp. 572-582.
Hawkins N. M., and D. Mitchell. (1979). Progressive Collapse of Flat-Plate Structures.
ACI Structural Journal, V. 76 No. 7, pp. 775-808.
Hawkins N. M., Bao A., and J. Yamazaki. (1989). Moment Transfer from Concrete Slabs
to Columns. ACI Structural Journal, V. 86, No. 6, pp. 705-716.

94
Hwang S. J. (1989). An Experimental Study of Flat-Plate Structures under Vertical and
Lateral Loads. PhD thesis, Department of Civil Engineering, University of California at
Berkeley.
Islam S., and R. Park. (1976). Tests on Slab-Column Connections with Shear and
Unbalanced Flexure. Journal of the Structural Division, ASCE, V. 102, No. ST3, Proc.
Paper 11972, pp 549-568.
Kanoh Y., and S. Yoshizaki. (1979). Strength of Slab-Column Connections Transferring
Shear and Moment. ACI Structural Journal, V. 76, No. 3, pp 461-478.
Kosut G. M., Burns M., and C. V. Winter. (1985). Test of Four-Panel Post-Tensioned
Flat Plate. Journal of the Structural Division, ASCE, V. 111, No. 8, pp 1916-1929.
Long A. E., and D. J. Cleland. (1993). Post-Tensioned Concrete Flat Slabs at Edge
Columns. Materials Journal of the American Concrete Institute, V. 90, No. M22, pp 207-
213.
Luo Y. H., Durrani A. J., and J. P. Conte. (1994). Equivalent Frame Analysis of Flat
Plate Building for Seismic Loading. Journal of Structural Engineering, ASCE, V. 120,
No. 7, pp 2137-2155.
Luo Y. H., and A. J. Durrani. (1995). Equivalent Beam Model for Flat-Slab Buildings –
Part I: Interior Connections. ACI Structural Journal, V. 92, No. 1, pp. 115-124.
Martinez-Cruzado J. A. (1993). Experimental Study of Post-tensioned Flat Plate Exterior
Slab-Column Connections Subjected to Gravity and Biaxial Loading. PhD thesis,
Department of Civil Engineering, University of California at Berkeley.
Mattock A. H. (1967). Discussion of “Rotational Capacity of Reinforced Concrete
Beams.” by Corley W. G., Journal of Structural Division, ASCE, V. 93, No. ST2, pp
512-522.
Megally S., and A. Ghali. (1994). Design Considerations for Slab-Column Connections
in Seismic Zones. ACI Structural Journal, V. 91, No. 30, pp 303-314.
Megally S., and A. Ghali. (2000). Punching Shear Design of Earthquake-Resistant Slab-
Column Connections. ACI Structural Journal, V. 97, No. 5, pp 720-730.
Moehle J. P., and J. W. Diebold. (1984). Experimental Study of the Seismic Response of a
Two-Story Flat-Plate Structure. Earthquake Engineering Research Center, Report No.
UCB/EERC-84/08, University of California at Berkeley.
Moehle J. P. (1998). Strength of Slab-Column Edge Connections. ACI Structural Journal,
V. 85, No. 11, pp 89-98.
Moehle J. P. (1996). Seismic Design Considerations for Flat-Plate Construction. ACI
SP-162: Mete A. Sozen Symposium, American Concrete Institute, Farmington Hills, MI,
pp. 1-34.

95
Mortin J. D., and A. Ghali. (1991). Connection of Flat Plates to Edge Columns. ACI
Structural Journal, V. 88, No. 2, pp 191-198.
Naaman A. E., and F. M. Alkhairi. (1991). Stress at Ultimate in Unbonded Post-
Tensioning Tendons: Part 1 – Evaluation of the state-of-the-art. ACI Structural Journal,
V. 88, No. 5, pp. 641-651.
Northridge Earthquake Reconnaissance Report, Vol. 2. (1996). Earthquake Spectra,
Supplement C to Volume 11, Earthquake Engineering Research Institute, Oakland, CA.
OpenSees Development Team. (2000). OpenSees: Open System for Earthquake
Engineering Simulations. Version 1.4, Berkeley, California.
Pan A. D., and J. P. Moehle. (1989). Lateral Displacement Ductility of Reinforced
Concrete Flat Plates. ACI Structural Journal, V. 86, No. 3., pp. 250-258.
Pan A. D., and J. P. Moehle. (1992). An Experimental Study of Slab-Column
Connections. ACI Structural Journal, V. 89, No. 6., pp. 626-638.
Parakash V., Powell G. H., and S. Campbell. (1993). DRAIN-2DX Base Program
Description and User Guide. Version 1.10, Structural Engineering Mechanics and
Materials, Report No. UCB/SEMM-93/17, University of California at Berkeley.
Park R., and W. L. Gamble. (2000). Reinforced Concrete Slabs 2ed. New York: John
Wiley & Sons.
Park R., Priestley M. J., and W. D. Gill. (1982). Ductility of Square-Confined Concrete
Columns. Journal of the Structural Division, ASCE, V. 108, No. ST4, pp 135-137.
Peabody D. Jr. (1948). Continuous Frame Analysis of Flat Slabs. Journal, Boston Society
of Civil Engineers.
Pecknold, D. A. (1975). Slab Effective Width for Equivalent Frame Analysis. ACI
Structural Journal, V.72, No. 4., pp. 135-137.
Pilakoutas, K., and C. Ioannou. (2000). Verification of a Novel Punching Shear
Reinforcement System of Flat Slabs. Procs International Workshop on Punching Shear
Capacity of RC Slabs, TRITA-BKN Bulletin 57, Royal Institute of Technology,
Stockholm.
PTI Committee for Unbonded Tendons. (2000). Field Procedures Manual for Unbonded
Single Strand Tendons. Post-Tensioning Institute, Phoenix, AZ.
Robertson I. N., and A. J. Durrani. (1990). Seismic Response of Connections in
Indeterminate Flat-Slab Subssemblies. Report No.41, Department of Civil Engineering,
Rice University.
Robertson, I. N., T. Kawai, J. Lee, and B. Enomoto. (2002). Cyclic Testing of Slab-
Column Connections with Shear Reinforcement. ACI Structural Journal, V. 99, No. 5, pp.
605-613.

96
Qaisrani A.-N. (1993). Interior Post-Tensioned Flat-Plate Connections Subjected to
Vertical and Biaxial Lateral Loading. PhD thesis, Department of Civil Engineering,
University of California at Berkeley.
Sawyer H. A. (1964). Design of Concrete Frames for Two Failure States. Proceedings of
the International Symposium on the Flexural Mechanics of Reinforced Concrete, ASCE-
ACI, Miami, pp 405-431.
Smith S. W., and N. H. Burns. (1974). Post-Tensioned Flat Plate to Column Connection
Behavior. Journal of the Prestressed Concrete Institute, V .19, No.3, pp 75-91.
Trongtham N., and N. M. Hawkins. (1977). Moment Transfer to Columns in Unbonded
Post-Tensioned Prestressed Concrete Slabs. Report SM77-3, Department of Civil
Engineering, University of Washington at Seattle.
Uniform Building Code. (1955). International Conference of Building Officials, Whittier,
California.
Uniform Building Code. (1997). International Conference of Building Officials, Whittier,
California.
Vanderbilt M. D., and W. G. Corley. (1983). Frame Analysis of Concrete Building.
Concrete International: Design and Construction, V. 5, No. 12, pp. 33-43.
Wallace J. W. (1992). BIAX: Revision 1 - A Computer Program for the Analysis of
Reinforced Concrete Sections. Report No. CU/CEE-92/4, Clarkson University, 1992.
Zee H. L., and J. P. Moehle. (1984). Behavior of Interior and Exterior Flat Plate
Connections Subjected to Inelastic Load Reversals. Earthquake Engineering Research
Center, Report No. UCB/EERC-84/07, University of California at Berkeley.

97
98
Table 2-1. Ratios of measured and calculated shear strength
using eccentric shear stress model

Edge connection
Mark νmeasured/ νcalculated Mark νmeasured/ νcalculated
Trongtham
Burns and Slab II-1 0.86 S2 0.82
and Hawkins
Hemakom
Slab II-2 0.83 S1 1.23
S2 1.21 S2 1.03
Kosut, Burns, Sunidja et. al
S6 1.33 S3 1.12
and Winter
S8 1.12 S4 1.12
E1 1.62 Martinez E1 1.91/0.84*
E2 1.62 E2 1.34/0.93*
Long and
E3 1.72
Cleland
E4 1.52
E5 1.88
Interior connection
Mark νmeasured/ νcalculated Mark νmeasured/ νcalculated
S1 0.90 Slab I-5 1.04
Smith and
S2 0.99 Slab I-6 1.35
Burns
S3 1.05 Slab I-7 1.25
Kosut, Burns,
5 1.12 Slab I-9 0.94
and Winter
S1 1.03 Burns and Slab I-10 1.37
Hemakom
S3 1.04 Slab I-11 1.04
Trongtham
S4 0.98 Slab I-15 0.96
and Hawkins
S5 1.09 Slab II-9 0.76
S6 0.90 Slab II-10 0.95
*
1 1.01/1.01 Slab II-15 0.68
*
Qaisrani 2 1.06/1.12
3 1.51/1.53*

* Ratios in two transverse directions subjected to biaxial loading

99
Table 2-2. Test variables and failure modes (Dilger and Shatila, 1989)

Test Variables Failure Modes


No Shear
S1 No Overhang Yielding Followed by Punching
Reinforcement
With Shear
S2 No Overhang Yielding Followed by Punching
Reinforcement
With Shear
S3 No Overhang Yielding Followed by Punching
Reinforcement
With Shear Flexural Failure followed by
S4 No Overhang
Reinforcement Punching
No Shear
S5 With Overhang Punching
Reinforcement
With Shear Flexural Failure followed by
S6 With Overhang
Reinforcement Punching

Table 2-3. Test and response summary (Moehle and Diebold, 1984)

Peak Max. Max. Max. Max. Free


Spectrum Damp.
Test Base Top Top Base Response Vib.
Intensity ratio
# Accel. Disp. Accel. Shear Period Period
Ratio (%)
(g) (mm) (g) (kN) (sec) (sec)

EQ1 0.015 0.03 - 0.025 4.1 - 0.21 1.5

EQ2 0.012 (0.005) 0.03 - 0.024 4.1 - 0.22 1.1

EQ3 0.047 0.10 - 0.090 16.1 - 0.21 1.3

EQ4 0.048 0.10 1.1 0.090 15.5 0.23 0.21 1.5

EQ5 0.092 0.21 2.1 0.16 30.6 0.24 0.21 1.4

EQ6 0.189 0.44 5.1 0.35 61.9 0.26 0.21 2.5

EQ7 0.202 (0.042) 0.44 9.9 0.49 93.3 0.28 0.23 2.3

EQ8 0.284 0.89 20.4 0.73 137 0.34 0.24 2.7

EQ9 0.252 (0.106) 0.89 29.7 0.83 148 0.39 0.27 4.9

EQ10 0.606 1.96 61.9 1.04 165 0.54 0.31 4.9

EQ11 0.827 (0.197) 2.71 95.5 1.08 175 0.58 0.45 7.1

1 00
Table 3-1. Concrete compressive strength results of core samples

Sample Sample Recorded Length Correction Corrected


Compressive To Factor Compressive
ID Location Stress (MPa)
Stress (MPa) Depth (ASTM C-42)
RC-1-1 1st Floor Slab 27.87 1.38 0.95 26.48
RC-1-2 1st Floor Slab 26.40 1.43 0.95 25.10
st
RC-1-3 1 Floor Slab 29.94 1.46 0.95 28.48
Average strength : 26.68
nd
RC-2-1 2 Floor Slab 19.06 1.43 0.95 18.13
nd
RC-2-2 2 Floor Slab 19.80 1.41 0.95 18.82
RC-2-3 2nd Floor Slab 22.49 1.40 0.95 21.37
nd
RC-2-4 2 Floor Slab 23.36 1.48 0.96 22.41
RC-2-5 2nd Floor Slab 17.22 1.46 0.95 16.34
nd
RC-2-6 2 Floor Slab 21.55 1.45 0.95 20.48
Average strength : 19.60
st
PT-1-1 1 Floor Slab 23.36 1.32 0.94 22.00
st
PT-1-2 1 Floor Slab 18.31 1.29 0.93 17.03
PT-1-3 1st Floor Slab 18.55 1.31 0.94 17.44
Average strength : 18.82
PT-2-1 2nd Floor Slab 24.33 1.29 0.93 22.62
nd
PT-2-2 2 Floor Slab 21.92 1.26 0.93 20.41
nd
PT-2-3 2 Floor Slab 19.99 1.30 0.94 18.82
Average strength : 20.62

Table 3-2. Reinforcing steel ratios for various conditions

c c + 3h Column strip Middle strip Full span

Width (b) 203 mm 470 mm 1029 mm 1029 mm 2057 mm

Top bars 1.38 1.38 1.00 0.45 0.72


Bottom bars 0.92 0.79 0.63 0.45 0.54
All – [%], ρ = As/(bd), d = 76.2 mm

101
Table 3-3. Concrete compressive strength results of cylinders

Sample ID Age Dimension (mm) Fracture Compressive Stress


Type* (MPa)
1 batch-1 7 152.4 × 304.8 C 14.07
1 batch-2 7 152.4 × 304.8 D 14.20
1 batch-3 14 152.4 × 304.8 D 18.27
1 batch-4 14 152.4 × 304.8 C 17.51
1 batch-5 28 152.4 × 304.8 D 19.93
1 batch-6 28 152.4 × 304.8 D 20.00
1 batch-7 28 152.4 × 304.8 C 20.75
1 bacth-8 35 152.4 × 304.8 D 21.51
1 bacth-9 35 152.4 × 304.8 D 21.37
2 batch-1 7 152.4 × 304.8 C 18.62
2 batch-2 7 152.4 × 304.8 C 17.93
2 batch-3 14 152.4 × 304.8 C 23.51
2 batch-4 14 152.4 × 304.8 C 24.20
2 batch-5 28 152.4 × 304.8 C 29.44
2 batch-6 28 152.4 × 304.8 C 29.37
2 batch-7 28 152.4 × 304.8 C 29.17

* Fracture Types : A=cone, B=cone and split, C=cone and shear, D=shear, E=columnar

1 02
Table 3-4. Material properties of steel bars and seven-wire strands

Diameter Area Yield Strength Tensile Strength


Sample ID (mm) (MPa) (MPa)
(mm2)
#2 – 1 4.09 13.16 450.9*/355.1** 662.6
* **
#2 – 2 4.09 13.16 493.7 /405.4 689.5
* **
#2 – 3 4.06 12.97 499.2 /359.9 699.8
* **
Average yield strength: 481.3 /373.5 (MPa)
#3 – 1 6.35 31.68 458.5*/417.1** 659.8
#3 – 2 6.30 31.16 465.4*/392.3** 650.9
#3 – 3 6.30 31.16 449.5*/392.3** 645.3
* **
Average yield strength: 481.3 /373.5 (MPa)
#4 – 1 9.04 64.19 490.9 697.1
#4 – 2 9.07 64.58 475.0 694.3
#4 – 2 9.04 64.19 507.4 701.2
Average yield strength: 491.1 (MPa)
Studrails – 1 3.99 12.52 630.9 732.2
Studrails – 2 4.01 12.65 632.9 746.0
Stdurails – 3 4.09 13.16 640.5 750.1
Average yield strength: 634.8 (MPa)
Strands – 1 7.95 37.23 1646.5 1840.2
Strands - 2 7.95 37.75 1603.0 1827.1
Average yield strength: 1624.8 (MPa)

* Yield Strength @ 0.2% Offset, ** Strength at 0.002 strains

1 03
Table 4-1. Channel list of RC specimen

Channel # Channel ID Description Gauge Type


1 H1o stroke Horizontal Disp. N-S direction (SE Actuator) Disp.
2 H2o stroke Horizontal Disp. E-W direction (NE Actuator) Disp.
3 H3o stroke Horizontal Disp. N-S direction (NW Actuator) Disp.
4 H4o stroke Horizontal Disp. E-W direction (SW Actuator) Disp.
5 V1o stroke Vertical displacement (SE Actuator) Disp.
6 V2o stroke Vertical displacement (NE Actuator) Disp.
7 V3o stroke Vertical displacement (NW Actuator) Disp.
8 V4o stroke Vertical displacement (SW Actuator) Disp.
9 H1 -2 acc Horizontal Accel. N-S direction (SE Actuator) Accelerometer
10 H3- 4 acc Horizontal Accel. E-W direction (NE Actuator) Accelerometer
11 H4- 1 acc Horizontal Accel. N-S direction (NW Actuator) Accelerometer
12 H2- 3 acc Horizontal Accel. E-W direction (SW Actuator) Accelerometer
13 1v acc Vertical Acceleration (SE Actuator) Accelerometer
14 2v acc Vertical Acceleration (NE Actuator) Accelerometer
15 3v acc Vertical Acceleration (NW Actuator) Accelerometer
16 4v acc Vertical Acceleration (SW Actuator) Accelerometer
17 Strain Gauge #1 Figure 4-17 120 ohm, single
18 Strain Gauge #2 Figure 4-17 120 ohm, single
19 Strain Gauge #3 Figure 4-18 120 ohm, single
20 DCDT #56 Figure 4-31 +/− 3’’, DCDT
21 Strain Gauge #5 Figure 4-18 120 ohm, single
22 Strain Gauge #6 Figure 4-18 120 ohm, single
23 Strain Gauge #7 Figure 4-18 120 ohm, single
24 Strain Gauge #8 Figure 4-17 120 ohm, single
25 Strain Gauge #9 Figure 4-16 120 ohm, single
26 Strain Gauge #10 Figure 4-17 120 ohm, single
27 Strain Gauge #11 Figure 4-17 120 ohm, single
28 Strain Gauge #12 Figure 4-16 120 ohm, single
29 Strain Gauge #13 Figure 4-16 120 ohm, single
30 Strain Gauge #14 Figure 4-16 120 ohm, single
31 Strain Gauge #15 Figure 4-16 120 ohm, single
32 Strain Gauge #16 Figure 4-16 120 ohm, single

104
33 Strain Gauge #17 Figure 4-16 120 ohm, single
34 Strain Gauge #18 Figure 4-16 120 ohm, single
35 Strain Gauge #19 Figure 4-17 120 ohm, single
36 Strain Gauge #20 Figure 4-16 120 ohm, single
37 Strain Gauge #21 Figure 4-18 120 ohm, single
38 Strain Gauge #22 Figure 4-18 120 ohm, single
39 DCDT #57 Figure 4-30 +/− 3’’, DCDT
40 Strain Gauge #24 Figure 4-18 120 ohm, single
41 Strain Gauge #25 Figure 4-18 120 ohm, single
42 Strain Gauge #26 Figure 4-16 120 ohm, single
43 Strain Gauge #27 Figure 4-17 120 ohm, single
44 Strain Gauge #28 Figure 4-17 120 ohm, single
45 Strain Gauge #29 Figure 4-16 120 ohm, single
46 Strain Gauge #30 Figure 4-17 120 ohm, single
47 Strain Gauge #31 Figure 4-17 120 ohm, single
48 Strain Gauge #32 Figure 4-16 120 ohm, single
49 Strain Gauge #33 Figure 4-17 120 ohm, single
50 Strain Gauge #34 Figure 4-16 120 ohm, single
51 Strain Gauge #35 Figure 4-16 120 ohm, single
52 Strain Gauge #36 Figure 4-17 120 ohm, single
53 Strain Gauge #37 Figure 4-16 120 ohm, single
54 Strain Gauge #38 Figure 4-16 120 ohm, single
55 Strain Gauge #39 Figure 4-18 120 ohm, single
56 Strain Gauge #40 Figure 4-18 120 ohm, single
57 Strain Gauge #41 Figure 4-17 120 ohm, single
58 Strain Gauge #42 Figure 4-16 120 ohm, single
59 Strain Gauge #43 Figure 4-17 120 ohm, single
60 Strain Gauge #44 Figure 4-16 120 ohm, single
61 Strain Gauge #45 Figure 4-18 120 ohm, single
62 Strain Gauge #46 Figure 4-18 120 ohm, single
63 Strain Gauge #47 Figure 4-18 120 ohm, single
64 Strain Gauge #48 Figure 4-18 120 ohm, single
65 VPM1, A NE footing, Axial force Tri-Axial, L.C.
66 VPM1, S, E-W NE footing, Shear, E-W direction Tri-Axial, L.C.
67 VPM1, S, N-S NE footing, Shear, N-S direction Tri-Axial, L.C.

105
68 VPM1, M, E-W NE footing, Moment, E-W direction Tri-Axial, L.C.
69 VPM1, M, N-S NE footing, Moment, N-S direction Tri-Axial, L.C.
70 VPM2, A NC footing, Axial force Tri-Axial, L.C.
71 VPM2, S, E-W NC footing, Shear, E-W direction Tri-Axial, L.C.
72 VPM2, S, N-S NC footing, Shear, N-S direction Tri-Axial, L.C.
73 VPM2, M, E-W NC footing, Moment, E-W direction Tri-Axial, L.C.
74 VPM2, M, N-S NC footing, Moment, N-S direction Tri-Axial, L.C.
75 VPM3, A NW footing, Axial force Tri-Axial, L.C.
76 VPM3, S, E-W NW footing, Shear, E-W direction Tri-Axial, L.C.
77 VPM3, S, N-S NW footing, Shear, N-S direction Tri-Axial, L.C.
78 VPM3, M, E-W NW footing, Moment, E-W direction Tri-Axial, L.C.
79 VPM3, M, N-S NW footing, Moment, N-S direction Tri-Axial, L.C.
80 VPM4, A SW footing, Axial force Tri-Axial, L.C.
81 VPM4, S, E-W SW footing, Shear, E-W direction Tri-Axial, L.C.
82 VPM4, S, N-S SW footing, Shear, N-S direction Tri-Axial, L.C.
83 VPM4, M, E-W SW footing, Moment, E-W direction Tri-Axial, L.C.
84 VPM4, M, N-S SW footing, Moment, N-S direction Tri-Axial, L.C.
85 VPM5, A SC footing, Axial force Tri-Axial, L.C.
86 VPM5, S, E-W SC footing, Shear, E-W direction Tri-Axial, L.C.
87 VPM5, S, N-S SC footing, Shear, N-S direction Tri-Axial, L.C.
88 VPM5, M, E-W SC footing, Moment, E-W direction Tri-Axial, L.C.
89 VPM5, M, N-S SC footing, Moment, N-S direction Tri-Axial, L.C.
90 VPM6, A SE footing, Axial force Tri-Axial, L.C.
91 VPM6, S, E-W SE footing, Shear, E-W direction Tri-Axial, L.C.
92 VPM6, S, N-S SE footing, Shear, N-S direction Tri-Axial, L.C.
93 VPM6, M, E-W SE footing, Moment, E-W direction Tri-Axial, L.C.
94 VPM6, M, N-S SE footing, Moment, N-S direction Tri-Axial, L.C.
95 Accel. #1 Figure 4-25 Accelerometer
96 Accel. #2 Figure 4-25 Accelerometer
97 Accel. #3 Figure 4-26 Accelerometer
98 Accel. #4 Figure 4-26 Accelerometer
99 Accel. #5 Figure 4-26 Accelerometer
100 Accel. #6 Figure 4-26 Accelerometer
101 Accel. #7 Figure 4-27 Accelerometer
102 Accel. #8 Figure 4-27 Accelerometer

106
103 Accel. #9 Figure 4-27 Accelerometer
104 Accel. #10 Figure 4-27 Accelerometer
105 DCDT #1 Figure 4-30 +/− 2’’, DCDT
106 DCDT #2 Figure 4-30 +/− 2’’, DCDT
107 DCDT #3 Figure 4-30 +/− 2’’, DCDT
108 DCDT #4 Figure 4-30 +/− 2’’, DCDT
109 DCDT #5 Figure 4-30 +/− 2’’, DCDT
110 DCDT #6 Figure 4-30 +/− 2’’, DCDT
111 DCDT #7 Figure 4-30 +/− 2’’, DCDT
112 DCDT #8 Figure 4-30 +/− 2’’, DCDT
113 DCDT #9 Figure 4-30 +/− 2’’, DCDT
114 DCDT #10 Figure 4-30 +/− 2’’, DCDT
115 DCDT #11 Figure 4-30 +/− 2’’, DCDT
116 DCDT #12 Figure 4-30 +/− 2’’, DCDT
117 DCDT #13 Figure 4-30 +/− 2’’, DCDT
118 DCDT #14 Figure 4-25 +/− 1’’, DCDT
119 DCDT #15 Figure 4-26 +/− 1’’, DCDT
120 DCDT #16 Figure 4-25 +/− 1’’, DCDT
121 DCDT #17 Figure 4-26 +/− 1’’, DCDT
122 DCDT #18 Figure 4-25 +/− 1’’, DCDT
123 DCDT #19 Figure 4-30 +/− 1’’, DCDT
124 DCDT #20 Figure 4-27 +/− 1’’, DCDT
125 W.P. #1 Figure 4-31 +/− 15’’, W.P.
126 W.P. #2 Figure 4-31 +/− 15’’, W.P.
127 Bad Channel
128 W.P. #4 Figure 4-31 +/− 15’’, W.P.
129 W.P. #5 Figure 4-30 +/− 15’’, W.P.
130 W.P. #6 Figure 4-31 +/− 15’’, W.P.
131 W.P. #7 Figure 4-31 +/− 15’’, W.P.
132 W.P. #8 Figure 4-31 +/− 15’’, W.P.
133 W.P. #9 Figure 4-31 +/− 15’’, W.P.
134 W.P. #10 Figure 4-31 +/− 15’’, W.P.
135 W.P. #11 Figure 4-30 +/− 15’’, W.P.
136 W.P. #12 Figure 4-30 +/− 15’’, W.P.
137 DCDT #25 Figure 4-27 +/− 1’’, DCDT

107
138 DCDT #26 Figure 4-27 +/− 1’’, DCDT
139 DCDT #27 Figure 4-27 +/− 1’’, DCDT
140 DCDT #28 Figure 4-27 +/− 1’’, DCDT
141 DCDT #29 Figure 4-27 +/− 1’’, DCDT
142 DCDT #30 Figure 4-27 +/− 1’’, DCDT
143 DCDT #31 Figure 4-27 +/− 1’’, DCDT
144 DCDT #32 Figure 4-27 +/− 1’’, DCDT
145 DCDT #33 Figure 4-26 +/− 1’’, DCDT
146 DCDT #34 Figure 4-26 +/− 1’’, DCDT
147 DCDT #35 Figure 4-27 +/− 1’’, DCDT
148 DCDT #36 Figure 4-30 +/− 1’’, DCDT
149 DCDT #37 Figure 4-26 +/− 1’’, DCDT
150 DCDT #38 Figure 4-26 +/− 1’’, DCDT
151 DCDT #39 Figure 4-27 +/− 1’’, DCDT
152 DCDT #40 Figure 4-26 +/− 2’’, DCDT
153 DCDT #41 Figure 4-26 +/− 0.5’’, DCDT
154 DCDT #42 Figure 4-27 +/− 0.5’’, DCDT
155 DCDT #43 Figure 4-25 +/− 0.5’’, DCDT
156 DCDT #44 Figure 4-25 +/− 0.5’’, DCDT
157 DCDT #45 Figure 4-25 +/− 0.5’’, DCDT
158 DCDT #46 Figure 4-25 +/− 0.5’’, DCDT
159 DCDT #47 Figure 4-25 +/− 0.5’’, DCDT
160 DCDT #48 Figure 4-27 +/− 0.5’’, DCDT
161 DCDT #49 Figure 4-27 +/− 0.5’’, DCDT
162 DCDT #50 Figure 4-25 +/− 0.5’’, DCDT
163 DCDT #21 Figure 4-27 +/− 1’’, DCDT
164 DCDT #22 Figure 4-27 +/− 1’’, DCDT
165 DCDT #23 Figure 4-25 +/− 1’’, DCDT
166 DCDT #24 Figure 4-26 +/− 1’’, DCDT
167 DCDT #51 Figure 4-25 +/− 1’’, DCDT
168 SP #12 Figure 4-30 +/− 2’’, SP
169 SP #13 Figure 4-30 +/− 2’’, SP
170 SP #14 Figure 4-30 +/− 2’’, SP
171 SP #15 Figure 4-30 +/− 2’’, SP
172 SP #16 Figure 4-30 +/− 2’’, SP

108
173 DCDT #52 Figure 4-30 +/− 3’’, DCDT
174 DCDT #53 Figure 4-31 +/− 3’’, DCDT
175 DCDT #54 Figure 4-31 +/− 3’’, DCDT
176 DCDT #55 Figure 4-30 +/− 3’’, DCDT
177 Pull back LC Tension force, Pull-back Test Tension Loadcell
178 SP #1 Figure 4-25 +/− 1’’, SP
179 SP #2 Figure 4-25 +/− 1’’, SP
180 SP #3 Figure 4-25 +/− 1’’, SP
181 SP #4 Figure 4-25 +/− 1’’, SP
182 SP #5 Figure 4-25 +/− 1’’, SP
183 SP #6 Figure 4-26 +/− 1’’, SP
184 SP #7 Figure 4-26 +/− 1’’, SP
185 SP #8 Figure 4-26 +/− 1’’, SP
186 SP #9 Figure 4-26 +/− 1’’, SP
187 SP #10 Figure 4-26 +/− 1’’, SP
188 SP #11 Figure 4-30 +/− 2’’, SP
189 W.P. #3 Figure 4-31 +/− 15’’, W.P.
190 SP #17 Figure 4-30 +/− 2’’, SP
191 SP #18 Figure 4-30 +/− 2’’, SP
192 SP #19 Figure 4-30 +/− 2’’, SP
193 SP #20 Figure 4-30 +/− 2’’, SP

109
Table 4-2. Channel list of PT specimen

Channel # Channel ID Description Gauge Type


1 H1o stroke Horizontal Disp. N-S direction (SE Actuator) Disp.
2 H2o stroke Horizontal Disp. E-W direction (NE Actuator) Disp.
3 H3o stroke Horizontal Disp. N-S direction (NW Actuator) Disp.
4 H4o stroke Horizontal Disp. E-W direction (SW Actuator) Disp.
5 V1o stroke Vertical displacement (SE Actuator) Disp.
6 V2o stroke Vertical displacement (NE Actuator) Disp.
7 V3o stroke Vertical displacement (NW Actuator) Disp.
8 V4o stroke Vertical displacement (SW Actuator) Disp.
9 H1 -2 acc Horizontal Accel. N-S direction (SE Actuator) Accelerometer
10 H3- 4 acc Horizontal Accel. E-W direction (NE Actuator) Accelerometer
11 H4- 1 acc Horizontal Accel. N-S direction (NW Actuator) Accelerometer
12 H2- 3 acc Horizontal Accel. E-W direction (SW Actuator) Accelerometer
13 1v acc Vertical Acceleration (SE Actuator) Accelerometer
14 2v acc Vertical Acceleration (NE Actuator) Accelerometer
15 3v acc Vertical Acceleration (NW Actuator) Accelerometer
16 4v acc Vertical Acceleration (SW Actuator) Accelerometer
17 VPM1, A NE footing, Axial force Tri-Axial, L.C.
18 VPM1, S, E-W NE footing, Shear, E-W direction Tri-Axial, L.C.
19 VPM1, M, E-W NE footing, Moment, E-W direction Tri-Axial, L.C.
20 VPM1, S, N-S NE footing, Shear, N-S direction Tri-Axial, L.C.
21 VPM1, M, N-S NE footing, Moment, N-S direction Tri-Axial, L.C.
22 VPM2, A NC footing, Axial force Tri-Axial, L.C.
23 VPM2, S, E-W NC footing, Shear, E-W direction Tri-Axial, L.C.
24 VPM2, M, E-W NC footing, Moment, E-W direction Tri-Axial, L.C.
25 VPM2, S, N-S NC footing, Shear, N-S direction Tri-Axial, L.C.
26 VPM2, M, N-S NC footing, Moment, N-S direction Tri-Axial, L.C.
27 VPM3, A NW footing, Axial force Tri-Axial, L.C.
28 VPM3, S, E-W NW footing, Shear, E-W direction Tri-Axial, L.C.
29 VPM3, M, E-W NW footing, Moment, E-W direction Tri-Axial, L.C.
30 VPM3, S, N-S NW footing, Shear, N-S direction Tri-Axial, L.C.
31 VPM3, M, N-S NW footing, Moment, N-S direction Tri-Axial, L.C.
32 VPM4, A SW footing, Axial force Tri-Axial, L.C.

110
33 VPM4, S, E-W SW footing, Shear, E-W direction Tri-Axial, L.C.
34 VPM4, M, E-W SW footing, Moment, E-W direction Tri-Axial, L.C.
35 VPM4, S, N-S SW footing, Shear, N-S direction Tri-Axial, L.C.
36 VPM4, M, N-S SW footing, Moment, N-S direction Tri-Axial, L.C.
37 VPM5, A SC footing, Axial force Tri-Axial, L.C.
38 VPM5, S, E-W SC footing, Shear, E-W direction Tri-Axial, L.C.
39 VPM5, M, E-W SC footing, Moment, E-W direction Tri-Axial, L.C.
40 VPM5, S, N-S SC footing, Shear, N-S direction Tri-Axial, L.C.
40 SP #8 Footing rotation, Run 4, 5 +/− 1’’, SP
41 VPM5, M, N-S SC footing, Moment, N-S direction Tri-Axial, L.C.
41 SP #9 Footing rotation, Run 4, 5 +/− 1’’, SP
42 VPM6, A SE footing, Axial force Tri-Axial, L.C.
43 VPM6, S, E-W SE footing, Shear, E-W direction Tri-Axial, L.C.
44 VPM6, M, E-W SE footing, Moment, E-W direction Tri-Axial, L.C.
45 VPM6, S, N-S SE footing, Shear, N-S direction Tri-Axial, L.C.
46 VPM6, M, N-S SE footing, Moment, N-S direction Tri-Axial, L.C.
47 Accel. #1 Figure 4-28 Accelerometer
48 Accel. #2 Figure 4-28 Accelerometer
49 Accel. #3 Figure 4-28 Accelerometer
50 Accel. #4 Figure 4-28 Accelerometer
51 Accel. #5 Figure 4-28 Accelerometer
52 Accel. #6 Figure 4-28 Accelerometer
53 Accel. #7 Figure 4-29 Accelerometer
54 Accel. #8 Figure 4-29 Accelerometer
55 Accel. #9 Figure 4-29 Accelerometer
56 Accel. #10 Figure 4-29 Accelerometer
57 Strain Gauge #1 Figure 4-20 120 ohm, single
58 Strain Gauge #2 Figure 4-20 120 ohm, single
59 Strain Gauge #3 Figure 4-20 120 ohm, single
60 Strain Gauge #4 Figure 4-20 120 ohm, single
61 Strain Gauge #5 Figure 4-19 120 ohm, single
62 Strain Gauge #6 Figure 4-19 120 ohm, single
63 Strain Gauge #7 Figure 4-20 120 ohm, single
64 Strain Gauge #8 Figure 4-20 120 ohm, single
65 Strain Gauge #9 Figure 4-20 120 ohm, single

111
66 Strain Gauge #10 Figure 4-20 120 ohm, single
67 Strain Gauge #11 Figure 4-19 120 ohm, single
68 Strain Gauge #12 Figure 4-19 120 ohm, single
69 Strain Gauge #13 Figure 4-20 120 ohm, single
70 Strain Gauge #14 Figure 4-20 120 ohm, single
71 Strain Gauge #15 Figure 4-19 120 ohm, single
72 Strain Gauge #16 Figure 4-19 120 ohm, single
73 Strain Gauge #17 Figure 4-19 120 ohm, single
74 Strain Gauge #18 Figure 4-19 120 ohm, single
75 Strain Gauge #19 Figure 4-20 120 ohm, single
76 Strain Gauge #20 Figure 4-20 120 ohm, single
77 Strain Gauge #21 Figure 4-19 120 ohm, single
78 Strain Gauge #22 Figure 4-19 120 ohm, single
79 Strain Gauge #23 Figure 4-19 120 ohm, single
80 Strain Gauge #24 Figure 4-19 120 ohm, single
81 Strain Gauge #25 Figure 4-19 120 ohm, single
82 Strain Gauge #26 Figure 4-19 120 ohm, single
83 Strain Gauge #27 Figure 4-19 120 ohm, single
84 Strain Gauge #28 Figure 4-19 120 ohm, single
85 Strain Gauge #29 Figure 4-19 120 ohm, single
86 Strain Gauge #30 Figure 4-19 120 ohm, single
87 Strain Gauge #31 Figure 4-19 120 ohm, single
88 Strain Gauge #32 Figure 4-19 120 ohm, single
89 Strain Gauge #33 Figure 4-19 120 ohm, single
90 Strain Gauge #34 Figure 4-19 120 ohm, single
91 Strain Gauge #35 Figure 4-20 120 ohm, single
92 Strain Gauge #36 Figure 4-19 120 ohm, single
93 Strain Gauge #37 Figure 4-19 120 ohm, single
94 Strain Gauge #38 Figure 4-19 120 ohm, single
95 Strain Gauge #39 Figure 4-20 120 ohm, single
96 Strain Gauge #40 Figure 4-19 120 ohm, single
97 Strain Gauge #41 Figure 4-20 120 ohm, single
98 Strain Gauge #42 Figure 4-20 120 ohm, single
99 Strain Gauge #43 Figure 4-19 120 ohm, single
100 Strain Gauge #44 Figure 4-20 120 ohm, single

112
101 Strain Gauge #45 Figure 4-20 120 ohm, single
102 Strain Gauge #46 Figure 4-20 120 ohm, single
103 Strain Gauge #47 Figure 4-20 120 ohm, single
104 Strain Gauge #48 Figure 4-20 120 ohm, single
105 Strain Gauge #49 Figure 4-20 120 ohm, single
106 Strain Gauge #50 Figure 4-19 120 ohm, single
107 Strain Gauge #51 Figure 4-19 120 ohm, single
108 Strain Gauge #52 Figure 4-19 120 ohm, single
109 Strain Gauge #53 Figure 4-19 120 ohm, single
110 Strain Gauge #54 Figure 4-20 120 ohm, single
111 Bad Channel
112 Strain Gauge #56 Figure 4-20 120 ohm, single
113 Strain Gauge #57 Figure 4-20 120 ohm, single
114 Strain Gauge #58 Figure 4-20 120 ohm, single
115 Strain Gauge #59 Figure 4-20 120 ohm, single
116 Strain Gauge #60 Figure 4-19 120 ohm, single
117 Strain Gauge #61 Figure 4-20 120 ohm, single
118 Strain Gauge #62 Figure 4-20 120 ohm, single
119 Strain Gauge #63 Figure 4-20 120 ohm, single
120 Strain Gauge #64 Figure 4-20 120 ohm, single
121 L.C. #1 FL1NW, Fig. 4-22 Comp. Loadcell
122 L.C. #2 FL1NW, Fig. 4-22 Comp. Loadcell
123 L.C. #3 FL1NW, Fig. 4-22 Comp. Loadcell
124 L.C. #4 FL1NE, Fig. 4-22 Comp. Loadcell
125 DCDT #1 Figure 4-32 +/− 3’’, DCDT
126 DCDT #2 Figure 4-32 +/− 3’’, DCDT
127 Strain Gauge #55 Figure 4-20 120 ohm, single
128 DCDT #3 Figure 4-33 +/− 3’’, DCDT
129 DCDT #4 Figure 4-32 +/− 3’’, DCDT
130 DCDT #5 Figure 4-33 +/− 3’’, DCDT
131 DCDT #6 Figure 4-33 +/− 3’’, DCDT
132 DCDT #7 Figure 4-32 +/− 2’’, DCDT
133 DCDT #8 Figure 4-33 +/− 2’’, DCDT
134 DCDT #9 Figure 4-32 +/− 2’’, DCDT
135 DCDT #10 Figure 4-32 +/− 2’’, DCDT

113
136 DCDT #11 Figure 4-32 +/− 2’’, DCDT
137 DCDT #12 Figure 4-32 +/− 2’’, DCDT
138 DCDT #13 Figure 4-32 +/− 2’’, DCDT
139 DCDT #14 Figure 4-32 +/− 2’’, DCDT
140 DCDT #15 Figure 4-32 +/− 2’’, DCDT
141 DCDT #16 Figure 4-33 +/− 2’’, DCDT
142 DCDT #17 Figure 4-28 +/− 1’’, DCDT
143 DCDT #18 Figure 4-28 +/− 1’’, DCDT
144 DCDT #19 Figure 4-28 +/− 1’’, DCDT
145 DCDT #20 Figure 4-29 +/− 1’’, DCDT
146 DCDT #21 Figure 4-29 +/− 1’’, DCDT
147 DCDT #22 Figure 4-28 +/− 1’’, DCDT
148 DCDT #23 Figure 4-28 +/− 1’’, DCDT
149 DCDT #24 Figure 4-28 +/− 1’’, DCDT
150 DCDT #25 Figure 4-29 +/− 1’’, DCDT
151 DCDT #26 Figure 4-28 +/− 1’’, DCDT
152 DCDT #27 Figure 4-29 +/− 1’’, DCDT
153 DCDT #28 Figure 4-29 +/− 1’’, DCDT
154 DCDT #29 Figure 4-28 +/− 1’’, DCDT
155 DCDT #30 Figure 4-29 +/− 1’’, DCDT
156 DCDT #31 Figure 4-29 +/− 1’’, DCDT
157 DCDT #32 Figure 4-29 +/− 1’’, DCDT
158 DCDT #33 Figure 4-28 +/− 1’’, DCDT
159 DCDT #34 Figure 4-28 +/− 1’’, DCDT
160 DCDT #35 Figure 4-29 +/− 1’’, DCDT
161 DCDT #36 Figure 4-29 +/− 1’’, DCDT
162 DCDT #37 Figure 4-29 +/− 1’’, DCDT
163 DCDT #38 Pull back, Free vib. test, Run 1, 2, 3, Fig. 4-32 +/− 1’’, DCDT
163 SP #7 Footing rotation, Run 4, 5 +/− 1’’, SP
164 DCDT #39 Figure 4-29 +/−0.5’’, DCDT
165 DCDT #40 Figure 4-28 +/−0.5’’, DCDT
166 DCDT #41 Figure 4-28 +/−0.5’’, DCDT
167 DCDT #42 Figure 4-28 +/−0.5’’, DCDT
168 DCDT #43 Figure 4-28 +/−0.5’’, DCDT
169 DCDT #44 Figure 4-28 +/−0.5’’, DCDT

114
170 DCDT #45 Figure 4-28 +/−0.5’’, DCDT
171 DCDT #46 Figure 4-28 +/−0.5’’, DCDT
172 DCDT #47 Figure 4-28 +/−0.5’’, DCDT
173 DCDT #48 Figure 4-28 +/−0.5’’, DCDT
174 DCDT #49 Figure 4-28 +/−0.5’’, DCDT
175 DCDT #50 Figure 4-28 +/−0.5’’, DCDT
176 DCDT #51 Figure 4-28 +/−0.5’’, DCDT
177 Pull back LC Tension force, Pell-back Test Tension Loadcell
178 W.P. #1 Figure 4-32 +/− 15’’, W.P.
179 W.P. #2 Figure 4-32 +/− 15’’, W.P.
180 W.P. #3 Figure 4-32 +/− 15’’, W.P.
181 W.P. #4 Figure 4-33 +/− 15’’, W.P.
182 W.P. #5 Figure 4-33 +/− 15’’, W.P.
183 W.P. #6 Figure 4-33 +/− 15’’, W.P.
184 W.P. #7 Figure 4-33 +/− 15’’, W.P.
185 W.P. #8 Figure 4-33 +/− 15’’, W.P.
186 W.P. #9 Figure 4-33 +/− 15’’, W.P.
187 W.P. #10 Figure 4-33 +/− 15’’, W.P.
188 SP #1 Footing Rotation +/− 1’’, SP
189 SP #2 Footing Rotation +/− 1’’, SP
190 SP #3 Footing Rotation +/− 1’’, SP
191 SP #4 Footing Rotation +/− 1’’, SP
192 SP #5 Footing Rotation +/− 1’’, SP
193 SP #6 Footing Rotation +/− 1’’, SP

Table 4-3. Characteristics of original ground motion

PGA PGV PGD Arias 5-75 % 5-95 % Mean


(g) (cm/sec) (cm) Intensity (sec) (sec) Period
0.135 10.24 7.18 57.30 cm/s 14.54 32.26 0.46 sec

115
Table 4-4. Free vibration test results (RC specimen)

Frequency (Hz) Damping ratio (%)


Before Test 4.32 1.33
After Run 1&2-1 3.53 3.05
After Run 2-2 3.39 3.31
After Run 3 3.07 4.51
After Run 4 2.02 6.74

Table 4-5. Free vibration test results (PT specimen)

Frequency (Hz) Damping ratio (%)


Before Test 3.61 1.86
After Run 1&2 3.24 2.65
After Run 3 2.96 3.59
After Run 4 2.26 4.67
After Run 5 2.13 5.04

116
Table 5-1. Various dimensions of displacement gauges (See Fig. 5-9)

RC specimen
Channel # r1 (mm) r2 (mm) h (mm)
105 54.0 155.6 28.6
106 50.8 152.4 28.6
107 57.2 142.9 34.9
108 54.0 177.8 77.8
109 46.0 155.6 22.2
110 46.0 154.0 25.4
111 50.8 163.5 30.2
112 55.6 163.5 36.5
113 49.2 136.5 38.1
114 54.0 144.5 30.2
115 47.6 155.6 22.2
116 46.0 144.5 20.6
117 44.5 139.7 25.4
118 22.2 119.1 98.4
119 15.9 136.5 11.1
120 30.2 127.0 98.4
121 22.2 134.9 30.2
122 27.0 122.2 68.3
124 25.4 146.1 73.0
137 19.1 108.0 68.3
138 11.1 138.1 65.1
139 6.4 136.5 77.8
140 27.0 115.9 73.0
141 20.6 112.7 65.1
142 12.7 128.6 71.4
143 15.9 138.1 60.3
144 12.7 131.8 65.1
145 28.6 152.4 77.8
146 38.1 127.0 22.2
147 28.6 130.2 79.4
148 44.5 154.0 30.2
149 14.3 133.4 71.4

117
150 41.3 155.6 82.6
151 30.2 171.5 69.9
152 44.5 149.2 69.9
153 31.8 136.5 71.4
154 1.6 104.8 68.3
155 31.8 128.6 74.6
156 47.6 127.0 73.0
157 38.1 119.1 30.2
158 50.8 146.1 73.0
159 25.4 127.0 74.6
160 20.6 125.4 69.9
161 38.1 155.1 65.1
162 25.4 139.7 73.0
163 22.2 122.2 79.4
164 25.4 127.0 71.4
165 22.2 108.0 68.3
166 27.0 128.6 58.7
167 22.2 130.2 79.4
168 69.9 161.9 50.8
169 73.0 152.4 50.8
170 63.5 155.6 50.8
171 50.8 158.8 60.3
172 57.2 161.9 54.0
173 63.5 155.6 50.8
174 54.0 149.2 44.5
175 60.3 149.2 50.8
176 47.6 133.4 47.5
193 38.1 130.2 63.5
194 54.0 108.0 60.2
195 47.6 117.5 60.3
196 47.6 136.5 57.2
197 19.1 136.5 60.3
198 34.9 136.5 31.8
199 25.4 139.7 57.2
200 44.5 136.5 73.0
201 47.6 127.0 63.5
202 57.2 108.0 50.8
203 34.9 136.5 60.3

118
PT specimen
Channel # r1 (mm) r2 (mm) h (mm)
132 63.5 123.8 7.9
134 58.7 119.1 6.4
136 66.7 155.6 28.6
137 63.5 152.4 23.8
138 60.3 122.2 14.3
140 58.7 122.2 19.1
142 50.8 128.6 60.3
143 57.2 144.5 81.0
144 34.9 146.1 77.8
145 74.6 141.3 69.9
146 23.8 112.7 73.0
147 60.3 144.5 68.3
148 33.3 131.8 76.2
149 66.7 152.4 66.7
150 22.2 114.3 69.9
151 125.4 201.6 65.1
152 31.8 117.5 63.5
153 36.5 122.2 69.9
154 61.9 123.8 63.5
155 108.0 203.2 65.1
156 38.1 114.3 73.0
157 39.7 125.4 76.2
158 42.9 146.1 142.9
159 38.1 150.8 69.9
160 31.8 127.0 61.9
161 71.4 150.8 69.9
162 34.9 114.3 69.9
164 33.3 122.2 69.9
165 73.0 158.8 73.0
166 63.5 130.2 66.7
167 66.7 144.5 54.0
168 38.1 177.8 88.9
169 66.7 138.1 25.4
170 103.2 165.1 63.5
171 44.5 130.2 69.9
172 34.9 131.8 66.7

119
173 34.9 131.8 66.7
174 71.4 139.7 60.3
175 44.5 117.5 71.4
176 38.1 114.3 78.2

Table 5-2. Drift capacity at punching of the RC specimen

Location φu,dcdt (r2)(φu,dcdt - φy) Vg/Vc θy,c+3h θu,c+3h

FL1NC-w -0.0040 -0.0144 -0.0130 -0.0274

FL1NC-e 0.0014 0.0021 0.0132 0.0153

FL1NC-avg 0.0082 0.26 0.0214

FL2NC-w -0.0030 -0.0074 -0.0130 -0.0204

FL2NC-e 0.0036 0.0137 0.0132 0.0270

FL2NC-avg 0.0105 0.32 0.0237

FL1SC-w N.A N.A N.A

FL1SC-e 0.0018 0.0038 0.26 0.0132 0.0170

FL1SC-avg N.A N.A N.A

FL1NE -0.0040 -0.0149 0.20 -0.0130 -0.0279

FL1NW N.A N.A N.A

FL2NE -0.0036 -0.0124 0.24 -0.0130 -0.0255

FL2NW 0.0060 0.0225 0.24 0.0132 0.0357

FL1SE -0.0040 -0.0160 0.20 -0.0130 -0.0290

FL1SW 0.0034 0.0102 0.20 0.0132 0.0234

120
Table 5-3. Drift capacity at punching of the PT specimen

Location φu,sg (r1)(φu,sg - φy) φu,dcdt (r2-r1)(φu,dcdt -φy) Vg/Vc θy,cs θu,cs

-0.0033 N.A N.A * -0.0261 -0.0379


FL1NC-w -0.0118

FL1NC-e 0.0040 0.0041 0.0025 0.0058 0.0205 0.0303

FL1NC-avg 0.35 0.0233 0.0341

FL2NC-w -0.0075 -0.0086 -0.0013 -0.0002 -0.0261 -0.0349

FL2NC-e 0.0048 0.0049 0.0015 0.0011 0.0204 0.0264

FL2NC-avg 0.0068 0.34 0.0233 0.0307**

FL1NE -0.0136 -0.0184 -0.0031 -0.0071 0.26 -0.0298 -0.0553

FL1NW 0.0028 0.0033 0.0021 0.0022 0.26 0.0223 0.0283

FL2NE -0.0127 -0.0149 -0.0021 -0.0027 0.25 -0.0298 -0.0474

FL2NW 0.0093 0.0229 0.0015 0.0005 0.25 0.0229 0.0463

*
: φu,sg is used for average curvature over the length r2.
**
: note that the punching capacity is at least this drift level.

121
Table 5-4. Post-tensioning

Description LC #1 LC #2 LC #3 LC #4
Loadcell location FL1NW-S FL1NW-C FL1NW-N FL1NE
End type Jacking end Dead end Jacking end Jacking end
(a pair w/ #4) (a pair w/ #2)
Tension Force [kN] First post-tensioning (2/20/02)
2/20/02, at jacking 26.20 13.74 33.76 19.26
2/20/02, after anchoring 19.04 13.74 25.98 12.05
2/20/02, after 3 hours 18.86 13.30 25.84 11.97
3/3/02, after 11 days 17.66 11.97 24.15 10.59
Tension Force [kN] Third post-tensioning (7/29/02)
7/29/02, at jacking, 47.24 48.17 50.89 51.73
7/29/02, after anchoring 39.86 46.22 41.64 42.66
7/29/02, after 20 minutes 39.63 45.86 41.64 42.66
7/29/02, after 2 hours 39.50 45.86 41.23 42.30
7/30/02, after 1 day 39.37 45.86 41.10 42.04

122
Table 5-5. Post-tensioning force change during testing

Description LC #1 LC #2 LC #3 LC #4
Loadcell location FL1NW-S FL1NW-C FL1NW-N FL1NE
End type Jacking end Dead end Jacking end Jacking end
(a pair w/ #4) (a pair w/ #2)
Tension Force [kN] RUN 1 (8/06/02)
8/06/02, before Run 1, 39.37 45.86 41.10 42.04
8/06/02, during Run 1, 39.32~39.41 45.82~45.91 41.01~41.28 41.99~42.21
8/06/02, after Run 1, 39.37 45.86 41.10 42.04
Tension Force [kN] RUN 2 (8/06/02)
8/06/02, during Run 2, 39.32~39.90 45.68~46.13 40.88~41.59 41.90~43.64
8/06/02, after Run 2, 39.46 45.68 42.66 42.66
Tension Force [kN] RUN 3 (8/08/02)
8/08/02, during Run 3, 39.37~41.86 44.62~46.48 40.83~43.50 42.57~45.46
8/08/02, after Run 3, 40.57 44.70 42.66 43.24
Tension Force [kN] RUN 4 (8/08/02)
8/08/02, during Run 4, 40.57~48.49 44.57~51.78 42.57~48.80 42.88~50.93
Max tendon force change
(∆f) 7.92 7.07 6.14 7.70

8/08/02, after Run 4, 41.72 44.75 43.73 44.17


Tension Force [kN] RUN 5 (8/08/02)
8/08/02, during Run 5, 41.59~49.29 44.53~52.04 43.55~49.64 43.44~51.11
Max tendon force change
(∆f) 7.56 7.30 5.92 6.94

8/08/02, after Run 5, 41.99 44.79 44.13 44.26

123
Table 6-1. Period comparisons for elastic element model

Fundamental Period (sec)

Specimen Model Free Simulated Ratio


Vibration Vibration T
( test )
Test Period Tsimulated

A 1.0 I col , α = 0.8, β = 1 0.214 1.32


B 1.0 I col , α = 0.8, β = 13 0.271 1.04

RC C 0.4 I col , α = 0.8, β = 13 0.290 0.98


0.283
Specimen D 0.7 I col ,α = 0.28 ~ 0.35, β = 1 0.267 1.06
E Fiber Col., α = 0.8, β = 1 0.217 1.30
F Fiber Col., α = 0.8, β = 13 0.272 1.04
A 1.0 I col , α = 0.65, β = 1 0.265 1.17
B 1.0 I col , α = 0.65, β = 1
2 0.310 1.0

PT C 0.4 I col , α = 0.65, β = 1


2 0.330 0.94
0.309
Specimen D 0.7 I col ,α = 0.28 ~ 0.35, β = 1 0.319 0.97
E Fiber Col., α = 0.65, β = 1 0.267 1.16
F Fiber Col., α = 0.65, β = 1
2 0.310 1.0

* Fundamental period was obtained prior to Run 2-2 (RC) and Run 2 (PT).

124
Table 6-2. Moment capacities of springs for each specimen

M y ,unb, f M y ,unb,v M y ,unb


Type or γf or or M n ,unb ∆M unb,v M u ,unb θu M fy ,cs
M ecc,unb, f M ecc,unb,v M ecc,unb

RC
Ext 7.57 1.0 0 7.57 8.81 3.28 12.09 0.028 13.33
(+)

RC
Ext 12.77 1.0 0 12.77 14.13 6.78 20.91 0.028 19.55
(–)

RC +13.33/
20.34 0.75 6.78 27.12 30.59 6.10 36.61 0.023
Int – 19.55

PT
Ext 6.90 0.617 4.29 11.19 · · 11.19 0.038* 14.24
(+)

PT
Ext 10.88 0.617 6.76 17.64 · · · 0.051 14.24
(–)

PT +7.57/
Int 21.70 0.6 14.46 36.16 · · · 0.032
– 22.37

* Elasto-perfectly plastic modeling until θu.


† All units for moments are kN-M.
†† Bold: governing values.

Munb Mu,unb Mslab


∑ M fy ,cs

Mn,unb γv = 1.0
My,unb Mfy,cs
Punching Yielding
Punching after Yielding within Column Strip,
prior to within c+3h Followed by
Yielding (item 2) Punching at θu - θ*
(item1) (item 3)

0 θ* θu θ 0 θ

125
126
Fig. 1-1 Slab - column connection with stud-rails (RC specimen)

c2+d

c1+d Vg γ v M u c AB
vdirect = vunb =
bo d Jc

(a) Slab - column critical section (b) Gravity shear (c) Eccentric shear
Fig. 1-2 Slab - column critical section and shear stress distributions

127
0.08
Conventional w/o S.R.
with Stud-rails
Post-tensioned

Rotation Capatity (Drift Ratio)


0.06

0.04 Ascheim & Moehle's


tentative model
for rotation capacity

0.02

0
0 0.2 0.4 0.6 0.8 1

Gravity Shear Ratio (Vg/φVc)

Fig. 1-3 Relationships between drift and gravity shear from experiments

“Stud-Rail”

Fig. 1-4 Punched connection with stud-rails superimposed on failure plan

128
Fig. 1-5 Test specimens

129
Fig. 2-1 Stiffness models for slab - column frame

Fig. 2-2 Effective beam width model proposed by Luo et al. (1994; 1995)

130
Fig. 2-3 Slab flexural yielding across the full width

Fig 2-4 Anchored flexural reinforcement within the transfer width

131
V = P2 - P1 V = P2 - P1
P1 M = M 1 + M2 P2 M = M 1 + M2
M1 M2
A A
c1 c2 c1 c2

C B c2+d C B c2 + d

c1+d c1+d/2
M2 M2
V M(c P2
P2
vAB = AC+ γV JC 1+d)
2
V
vAB = A + γV J
M cAB
V M V V C 2
vAB = AC- γV JC (c1+d)
C
V M cAB
2 vAB = AC- γV JC 2
M c2+d M c2 + d

(c1+d)/2 cCD cAB Centroid of


c1+d c1+d/2 Critical Section
Interior Column Edge Column

Fig 2-5 Shear and moment transfer assumed in eccentric shear stress model

Critical Sections

Critical Sections

Centorid of Critical Section

Interior Column Edge Column

Fig 2-6 Critical sections outside of shear-reinforced zone

132
Fig. 2-7 Ratio between measured and calculated moment transfer strengths
using the effective transfer width, c1 + 2c2 (Moehle, 1988)

0.07 Conventionally
0.06 Reinforced
Drift Ratio (% of Height)

0.05 Post-Tensioned

0.04
0.03
0.02
0.01
0
0 0.2 0.4 0.6 0.8 1
Gravity Shear Ratio

Fig. 2-8 Effect of gravity load on drift

133
Lateral Load Lateral Load
Transfer Points Shear Load Shear Load Transfer Points
Transfer Points Transfer Points

Slab Column

Shear Load Shear Load Lateral Load

Lateral Load

Fig. 2-9 Tests of isolated specimens conducted by Trongtham and Hawkins (1977)

Fig. 2-10 Comparison of lateral load-deflection relationships


(Hawkins et al., 1975; 1975; 1977)

134
Fig. 2-11 Moment-deflection relationships for four specimens (Foutch et al., 1990)

Fig. 2-12 Load vs. displacement envelopes (Qaisrani, 1993)

135
Fig. 2-13 Test structure on shake table (Moehle and Diebold, 1984)

Fig. 2-14 Envelope relationship between top displacement and base shear
(Moehle and Diebold, 1984)

136
Fig. 2-15 Computed and measured relations between top displacement and base shear
(Moehle and Diebold, 1984)

Fig. 2-16 Test structure (Hayes, Foutch, and Wood, 1999)

Fig. 2-17 Second-story drift comparison for El Centro simulations on bare model
(Hayes, Foutch, and Wood, 1999)

137
Fig. 3-1(a) RC specimen

Fig. 3-1(b) PT specimen

138
Fig. 3-2(a) Test specimens

Fig. 3-2(b) Prototype and shake table specimens (Hatched)

139
Fig. 3-3(a) Elevation view of RC specimen (Frame N, S)

Fig. 3-3(b) Elevation view of RC specimen (Frame W, C, E)

140
Fig. 3-4(a) Details of column

Fig. 3-4(b) Column section

141
Fig. 3-5(a) Details of footing

Fig. 3-5(b) Footing

142
Fig. 3-6(a) Top slab reinforcement of RC specimen

Fig. 3-6(b) Bottom slab reinforcement of RC specimen

143
Fig. 3-7(a) Details of top reinforcement (RC specimen, Exterior connection)

Fig. 3-7(b) Details of top reinforcement (RC specimen, Interior connections)

144
Fig. 3-8 Details of bottom reinforcement (RC specimen)

Fig. 3-9(a) Exterior connection (RC) Fig. 3-9(b) Interior connection (RC)

Fig. 3-9(c) Slab reinforcement of RC specimen

145
Column critical section– d/2 from column face

Outer critical section– d/2 from stud-rails

Fig. 3-10(a) Slab - column critical sections Fig. 3-10(b) Stud-rails

Ec = 2800 ksi (19305 MPa) 30


4000
fc' = 3773 psi (26 MPa)
25

Concrete Stress (MPa)


Concrete Stress (psi)

3000
20

2000 15

10
1000
5

0 0
0 0.001 0.002 0.003 0.004

Concrete Strain

Fig. 3-11 Stress - strain relations of first batch concrete cylinder

Fig. 3-12 Test specimens

146
Fig. 3-13(a) Elevation view of PT specimen (Frame N, S)

Fig. 3-13(b) Elevation view of PT specimen (Frame E, C, W)

147
Fig. 3-14(a) Tendon arrangement of PT specimen

Fig. 3-14(b) Plan views of bonded reinforcement (PT specimen)

148
Fig. 3-15(a) Overview of tendon arrangement (PT specimen)

Fig. 3-15(b) Interior connection (PT) Fig. 3-15(c) Edge connection (PT)

Fig. 3-16(a) Anchor plate and edge tension bar Fig. 3-16(b) Anchors

149
Fig. 3-17(a) Details of interior connection (PT specimen)

Fig. 3-17(b) Details of edge connection (PT specimen)

150
Fig. 3-18(a) Tendon layouts (E-W direction, See Fig. 3-12)

Fig. 3-18(b) Tendon layouts (N-W direction, See Fig. 3-12)

151
Fig. 3-19 Donut-shaped load cells

152
Fig. 4-1 Moving PT specimen with bracing system Fig. 4-2 Embedded rods

Fig. 4-3 Lifting anchor

Fig. 4-4 Baseplates for attachment of specimens

153
Fig. 4-5 Plans of baseplates

Steel Washer

Footing

406 mm x 406 mm
Steel Plate

Tri-axial Load cell

Baseplate

Fig. 4-6 Footing attachment

Fig. 4-7 Footing anchorage Fig. 4-8 Installation of lead-weights

154
Fig. 4-9 Layouts of lead-weights

Fig. 4-10 A pair of steel rollers

Rubber Pads
Steel Pads Rubber Pads

Fig. 4-11 Steel pads and rubber pads

155
Fig. 4-12 Wooden dowels Fig. 4-13 Re-shoring of PT slabs

Fig. 4-14(a) Strain gauges Fig. 4-14(b) Strain gauges applied with M-coat J-3

Fig. 4-15 Locations of strain gauges attached on slab reinforcement

156
Fig. 4-16 Strain gauges on top bars (RC specimen, FL1, FL2)

Fig. 4-17 Strain gauges on bottom bars (RC specimen, FL1, FL2)

157
Fig. 4-18 Strain gauges on column bars and stud-rails (RC specimen, Frame N)

Fig. 4-19 Strain gauges on slab bars and slab surface (PT specimen, FL1, FL2)

158
Fig. 4-20 Strain gauges on bars, stud-rails, and column surface (PT specimen, Frame N)

Fig. 4-21 Tri-axial load cell

159
Fig. 4-22 Locations of donut shaped load cells

Fig. 4-23 Donut-shaped load cells for tendons Fig. 4-24 Accelerometers

160
Fig. 4-25 Instrumentation on slab bottom and table (RC specimen, FL1)

Fig. 4-26 Instrumentation on slab top (RC specimen, FL1)

161
Fig. 4-27 Instrumentation on slab top and bottom (RC specimen, FL2)

Fig. 4-28 Instrumentation on slab top and bottom (PT specimen, FL1)

162
Fig. 4-29 Instrumentation on slab top and bottom (PT specimen, FL2)

Fig. 4-30 Instrumentation (RC specimen, Frame N)

163
Fig. 4-31 Instrumentation (RC specimen, Frame S)

Fig. 4-32 Instrumentation (PT specimen, Frame N)

164
Fig. 4-33 Instrumentation (PT specimen, Fame S)

Fig. 4-34 DCDTs mounted on top and bottom of slab

Fig. 4-35 DCDTs mounted on each side of column

165
Steel washer Threaded
Hook rods
Wire
Pulley

DCDT Core

DCDT Aluminium bar

Fig. 4-36 DC displacement transducer

Pulley Aluminium bar

Wire Hook SP

Steel washer
Additional spring

Fig. 4-37 Spring potentiometer

166
SP

SP

Fig. 4-38 SPs installed between footing and steel plate

Spring
Potentiometer 2

Spring
Potentiometer 3
Spring
Potentiometer 1

Deformed Footing Plane

Fig. 4-39(a) Rotation of footing plane monitored using SPs

Fig. 4-39(b) Locations of SPs monitoring footing rotation (PT specimen)

167
0.15

0.1
RC-RUN 1

Acceleration (g)
0.05

-0.05

-0.1

-0.15
0 5 10 15 20 25 30 35 40 45
Time (sec)

0.3

0.2
RC-RUN 2-2
Acceleration (g)

0.1

-0.1

-0.2

-0.3
0 5 10 15 20 25 30 35 40 45
Time (sec)

0.5
0.4 RC-RUN 3
0.3
Acceleration (g)

0.2
0.1
0
-0.1
-0.2
-0.3
-0.4
-0.5
0 5 10 15 20 25 30 35 40 45
Time (sec)

1.5
1.2 RC-RUN 4
0.9
Acceleration (g)

0.6
0.3
0
-0.3
-0.6
-0.9
-1.2
-1.5
0 5 10 15 20 25 30 35 40 45
Time (sec)

Fig. 4-40. Table acceleration histories of each test (RC specimen)

168
0.15

0.1
PT-RUN 1

Acceleration (g)
0.05

-0.05

-0.1

-0.15
0 5 10 15 20 25 30 35 40 45
Time (sec)
0.3

0.2
PT-RUN 2
Acceleration (g)

0.1

-0.1

-0.2

-0.3
0 5 10 15 20 25 30 35 40 45
Time (sec)
0.5
0.4 PT-RUN 3
0.3
Acceleration (g)

0.2
0.1
0
-0.1
-0.2
-0.3
-0.4
-0.5
0 5 10 15 20 25 30 35 40 45
Time (sec)
1.5
1.2 PT-RUN 4
0.9
Acceleration (g)

0.6
0.3
0
-0.3
-0.6
-0.9
-1.2
-1.5
0 5 10 15 20 25 30 35 40 45
Time (sec)
1.5
1.2 PT-RUN 5
0.9
Acceleration (g)

0.6
0.3
0
-0.3
-0.6
-0.9
-1.2
-1.5
0 5 10 15 20 25 30 35 40 45
Time (sec)

Fig. 4-41 Table acceleration histories of each test (PT specimen)

169
0.6 1.5

RC-RUN 1 RC-RUN 2-2


Spectral Acceleration (g) 0.5 1.25

Spectral Acceleration (g)


0.4 1

0.3 0.75 2% damping

2% damping
0.2 0.5

0.1 0.25
5% damping 5% damping
0 0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Period (sec) Period (sec)

2.4 7.2

2.1 RC-RUN 3 RC-RUN 4


6
Spectral Acceleration (g)

1.8

Spectral Acceleration (g)


4.8
1.5

1.2 2% damping 2% damping


3.6

0.9
2.4
0.6
1.2
0.3 5% damping 5% damping
0 0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Period (sec) Period (sec)
0.6 1.5

PT-RUN 1 PT-RUN 2
0.5 1.25
Spectral Acceleration (g)

Spectral Acceleration (g)

0.4 1

0.3 0.75 2% damping

2% damping
0.2 0.5

0.1 0.25
5% damping 5% damping
0 0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Period (sec) Period (sec)
2.4 7.2

2.1 PT-RUN 3 PT-RUN 4


6
Spectral Acceleration (g)
Spectral Acceleration (g)

1.8
4.8
1.5

1.2 2% damping 3.6 2% damping

0.9
2.4
0.6
1.2
0.3 5% damping 5% damping

0 0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Period (sec) Period (sec)

170
8.4

PT-RUN 5
Spectral Acceleration (g) 7

5.6

4.2 2% damping

2.8

1.4
5% damping

0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Period (sec)

Fig. 4-42 Response spectra of each test

Fig. 4-43 Time-compressed ground motion

171
δ2 z Footing Rotation
δ −δ
θ = 1b 2b
δ1 l
h z Top Relative Displacement
∆ = (δ − δ ) − h θ
δ0 2 0

δ1b δ2b z Base Shear – one frame

∑ V LC = V BaseShear
l
Fig. 5-1 Process used to compute base shear and top relative displacement

800
PT-RUN5 80
600
60

Moment, LC (kN-m)
Interior Footing
Moment, LC (in.-k)

400 40
200 20
Exterior Footing
0 0

-200 -20

-400 -40
-60
-600
Linear Fit -80
-800
-0.02 -0.015 -0.01 -0.005 0 0.005 0.01 0.015 0.02
Footing Rotation (rad)
Fig. 5-2 Footing rotation vs moment relations

172
Top Drift [%]
-3 -2 -1 0 1 2 3
RC-RUN2-2
60
RC-RUN4 1.5

40
1
Base Shear [kips]

Base Shear [W]


20 0.5

0 0

-20 -0.5

-1
-40

-1.5
-60

-3 -2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5 3

Top Relative Displacement [in.]

Fig. 5-3(a) Base shear vs top relative displacement relations (RC specimen)

Top Drift [%]


-5 -4 -3 -2 -1 0 1 2 3 4 5

60 PT-RUN2 1.5
PT-RUN4
PT-RUN5
40 1
Base Shear [kips]

Base Shear [W]

20 0.5

0 0

-20 -0.5

-40 -1

-60 -1.5

-4 -3 -2 -1 0 1 2 3 4

Top Relative Displacement [in.]

Fig. 5-3(b) Base shear vs top relative displacement relations (PT specimen)

173
400 400

Base Column Moment (kN-m)

Base Column Moment (kN-m)


Base Column Moment (in.-k)

Base Column Moment (in.-k)


RC-RUN4-FL0NE φ+y,col 40 RC-RUN4-FL0NW φ+y,col 40
300 300
30 30
200 20 200 20
100 10 100 10
0 0 0 0

-100 -10 -100 -10

-200 -20 -20


-200
-30 -30
-300 -300
φ−y,col -40 φ−y,col -40
-400 -400
-0.004 -0.003 -0.002 -0.001 0 0.001 0.002 0.003 0.004 -0.004 -0.003 -0.002 -0.001 0 0.001 0.002 0.003 0.004
Curvature (1/in.) Curvature (1/in.)

500 400

Base Column Moment (kN-m)


Base Column Moment (kN-m)
Base Column Moment (in.-k)

Base Column Moment (in.-k)


PT-RUN4-FL0NC φ+y,col 50 PT-RUN4-FL0NE φ+y,col 40
400 300
40 30
300
30 200
200 20
20
100 100 10
10
0 0 0 0
-100 -10 -10
-100
-200 -20
-20
-30 -200
-300
-40 -30
-400 -300
φ−y,col -50 φ−y,col -40
-500 -400
-0.004 -0.003 -0.002 -0.001 0 0.001 0.002 0.003 0.004 -0.004 -0.003 -0.002 -0.001 0 0.001 0.002 0.003 0.004
Curvature (1/in.) Curvature (1/in.)

Fig. 5-4 Moment vs curvature relations at the column base

Fig. 5-5 Damage at the column base (RC specimen)

174
Column Curvature (1/in.) 0.005
0.004 0.0016

Column Curvature (1/cm.)


0.003 0.0012
0.002 0.0008
0.001 0.0004
0 0
-0.001 -0.0004
-0.002 -0.0008
-0.003 -0.0012
-0.004
-0.005
CURVATURE (φ, Strain Gauges) -0.0016

8 9 10 11 12 13 14 15 16 17 18
Time (sec)

Curvature (φ,1/cm)
-0.00015 -7.5E-005 0 7.5E-005 0.00015
50
400
40
300 500
30 450 50

Moment (in.-k)
200 400
Moment (in.-k)

20

Moment (kN-m)
Moment (kN-m)
350 40
100 10 300
30
250 Column Yield Moment
0 0 200
20
150
-10
-100 100
50
MOMENT (M) 10

-200 -20 0 0
-30 12.62 12.63 12.64 12.65 12.66 12.67 12.68 12.69 12.7
-300 BIAX Time (sec)
-40
-400
-50
-0.004 -0.002 0 0.002 0.004
Curvature (φ,1/in.)

Fig. 5-6 Process used to compute moment from measured curvature

175
500
450 50
400

Moment (kN-m)
40

Moment (in.-k)
350
300
250 30
Column Yield Moment
200
MLOAD CELL 20
150
100 MDISP. GAUGE
10
50 RC-RUN4-FL0NC MSTRAIN GAUGE
0 0
12.62 12.63 12.64 12.65 12.66 12.67 12.68 12.69 12.7
Time (sec)

500
450 50
400

Moment (kN-m)
40
Moment (in.-k)

350
300
250 30
Column Yield Moment
200
20
150
M LOAD CELL
100 10
M STRAIN GAUGE
50 RC-RUN4-FL0NE
0 0
12.62 12.63 12.64 12.65 12.66 12.67 12.68 12.69 12.7
Time (sec)
500
450 50
400

Moment (kN-m)
40
Moment (in.-k)

350
300
250 30
Column Yield Moment
200
20
150 MLOAD CELL
100 MDISP. GAUGE 10
50 RC-RUN4-FL0NW MSTRAIN GAUGE
0 0
12.62 12.63 12.64 12.65 12.66 12.67 12.68 12.69 12.7
Time (sec)
140

120 14
12
Moment (kN-m)

100
Moment (in.-k)

10
80
8
60
6
40
4
MLOAD CELL
20 MDISP. GAUGE 2
RC-RUN4-FL1NC
0 0
12.62 12.63 12.64 12.65 12.66 12.67 12.68 12.69
Time (sec)

Fig. 5-7 Column moments from independent measures

176
0.005
0.004 0.0016

Column Curvature (1/cm.)


Column Curvature (1/in.)
0.003 0.0012
0.002 0.0008
0.001 0.0004
0 0
-0.001 -0.0004
-0.002 φLOAD CELL -0.0008
-0.003 φDISP. GAUGE -0.0012
-0.004 PT-RUN4-FL0NC φSTRAIN GAUGE -0.0016
-0.005
8 9 10 11 12 13 14 15 16 17 18
Time (sec)

Fig. 5-8 Column curvatures from independent measures

r1 = ~ 1.5 in. (38.1 mm)

h Displacement Gauge

~ 1 in. (25.4mm)
Strain Gauge
h

r2 = ~ 5 in. (127 mm)


Fig. 5-9 Strain gauges and displacement gauges used to measure slab curvature

177
1200
800 RC-RUN 4

Micro Strain (µs)


400
0
-400
-800
Top Bars
-1200
10 12 14 16 18 20
Time (sec)

Fig. 5-10(a) Slab strain gauge data located symmetrically (RC, FL1NW, Top bars)
2000
1500 RC-RUN 4
Micro Strain (µs)

1000
500
0
-500
-1000
-1500 Bottom Bars
-2000
10 12 14 16 18 20
Time (sec)

Fig. 5-10(b) Slab strain gauge data located symmetrically (RC, FL1NW, Bottom bars)

Distance from column center (cm)


-100 -80 -60 -40 -20 0
0.004 0.0015
0.003 FL1NW
Curvature (1/in.)

0.001
Curvature (1/cm)

0.002
φy+,c+3h 0.0005
0.001
0 0
-0.001 12.69 sec -0.0005
-0.002 11.40 sec φy-,c+3h
12.92 sec -0.001
-0.003 12.53 sec RC-RUN 4
-0.004 -0.0015
-40 -30 -20 -10 0
Distance from column center (in.)

Fig. 5-11 Slab curvature variations of exterior connections (RC, FL1NC-w)

178
p ( ) ( )
22000 6000
Effective Transfer Width
5000
18000

Bottom Bars Strain [µs]


damaged
Top Bars Strain [µs]
Top Bar 1 4000 3 2
14000
3000
Effective
Steel Yielding
Transfer Width
10000 2000

3 2 1 1000 Top Bar 2


6000
0 Top Bar 3
Steel Yielding
Top Bar 2
2000
-1000
Top Bar 3
-2000 -2000
5 10 15 20 25 30 35 40 45 5 10 15 20 25 30 35 40 45

Time [sec.] Time [sec.]

(a) Top bars (FL1NW) (b) Bottom bars (FL1NW)


Fig. 5-12 Slab strain gauge data (RC specimen)

8000 Top Bar 2


Top Bars Strain [µs]

6000

1 2 3 4
4000
C
2000 Steel Yielding
Top Bar 3
Top Bar 1
0 Top Bar 4

RC-RUN4-FL1NC-w
-2000
10 12 14 16 18 20
Time [sec.]

Fig. 5-13(a) Slab strain gauge data (RC, FL1NC, Top bars)

Bottom Bar 2

Steel Yielding
Bottom Bars Strain [µs]

2000

Bottom Bar 3

Bottom Bar 1
0

RC-RUN4-FL1NC-w
-2000
10 12 14 16 18 20
Time [sec.]

Fig. 5-13(b) Slab strain gauge data (RC, FL1NC, Bottom Bars)

179
Fig. 5-14(a) Damage on top slab at the end of testing (RC, FL1SE)

Fig. 5-14(b) Damage on bottom slab at the end of testing (RC, FL2SW)

(a) Flexural cracks (PT, FL2W-bottom) (b) Flexural cracks (PT, FL2C-top)
Fig. 5-15 Flexural cracks across the full width at the end of testing

180
Fig. 5-16(a) Damage on top slab at the end of testing (RC specimen, FL1SC)

Fig. 5-16(b) Damage on bottom slab at the end of testing (RC specimen, FL1NC)

(a) Slab top (FL2SC) (b) Slab bottom (FL2SC)


Fig. 5-17 Damage on top and bottom slabs at the end of testing (PT specimen)

181
Fig. 5-18 Torsional cracks on top slab at the end of testing (PT specimen, FL2SE)

Fig. 5-19 Substantial concrete spalling on west edge at the end of testing (RC, FL1SW)

182
Slab Curvature (1/cm)
-0.0003 -0.00015 0 0.00015 0.0003
0.0025
0.0006
φy+,col
0.0005 0.002
φy+,c+3h
0.0004 0.0015

Column Curvature (1/cm)


Column Curvature (1/in.) 0.0003
0.001
0.0002
0.0001 0.0005
0 0
-0.0001 -0.0005
-0.0002
-0.001
-0.0003
-0.0004 -0.0015
φy-,c+3h
-0.0005
φy-,col RC-RUN4-FL2NW -0.002
-0.0006
-0.0025
-0.008 -0.004 0 0.004 0.008
Slab Curvature (1/in.)

Fig. 5-20(a) Curvature diagrams (RC specimen, FL2NW)

Slab Curvature (1/cm)


-0.0003 -0.00015 0 0.00015 0.0003
0.0025
0.0006
+
φy ,col 0.002
0.0005 φy+,c+3h
0.0004 0.0015
Column Curvature (1/cm)
Column Curvature (1/in.)

0.0003
0.001
0.0002
0.0001 0.0005
0 0
-0.0001 -0.0005
-0.0002
-0.001
-0.0003
-0.0004 -0.0015
φy-,c+3h
-0.0005 φy-,col RC-RUN4-FL2NE -0.002
-0.0006
-0.0025
-0.008 -0.004 0 0.004 0.008
Slab Curvature (1/in.)

Fig. 5-20(b) Curvature diagrams (RC specimen, FL2NE)

183
Slab Curvature (1/cm)
-0.003 -0.0015 0 0.0015 0.003
0.00025
0.0006
PT-RUN4-FL2NC-w φy+,col
0.0005 φy+,c+3h 0.0002
0.0004 0.00015

Column Curvature (1/cm)


Column Curvature (1/in.)
0.0003
0.0001
0.0002
0.0001 5E-005
0 0
-0.0001 -5E-005
-0.0002
-0.0001
-0.0003
-0.0004 -0.00015
-0.0005 φy-,c+3h -0.0002
φy-,col
-0.0006
-0.00025
-0.009-0.006-0.003 0 0.003 0.006 0.009
Slab Curvature (1/in.)

Fig. 5-21(a) Curvature diagrams (PT specimen, FL2NC-w)

Slab Curvature (1/cm)


-0.003 -0.0015 0 0.0015 0.003
0.0025
0.0006
φy+,col 0.002
0.0005
0.0004 φy+,c+3h
0.0015
Column Curvature (1/cm)
Column Curvature (1/in.)

0.0003
0.001
0.0002
0.0001 0.0005
0 0
-0.0001 -0.0005
-0.0002
-0.001
-0.0003
-0.0004 -0.0015
-0.0005 φy-,c+3h
PT-RUN4-FL2NC-e -0.002
φy-,col
-0.0006
-0.0025
-0.009-0.006-0.003 0 0.003 0.006 0.009
Slab Curvature (1/in.)

Fig. 5-21(b) Curvature diagrams (PT specimen, FL2NC-e)

184
-0.003 -0.0015 0 0.0015 0.003
60
500 φy+,col
PT-RUN4-FL1NC-w 50
400
φy ,c+3h
+ 40

Unbalanced Moment (kN-m)


Unbalanced Moment (in.-k)
300
30
200 20
100 10
0 0
-100 -10

-200 -20
-30
-300
-40
-400 φy-,c+3h
-50
-500 φy-,col
-60
-0.009-0.006-0.003 0 0.003 0.006 0.009
Slab Curvature (1/in.)

Fig. 5-22(a) Curvature diagrams (PT specimen, FL1NC-w)

Slab Curvature (1/cm)


-0.003 -0.0015 0 0.0015 0.003
60
500 φy+,col
PT-RUN4-FL1NC-e 50
400 φy+,c+3h
40
Unbalanced Moment (kN-m)
Unbalanced Moment (in.-k)

300
30
200 20
100 10
0 0
-100 -10

-200 -20
-30
-300
φy-,c+3h -40
-400
-50
-500 φy-,col
-60
-0.009-0.006-0.003 0 0.003 0.006 0.009
Slab Curvature (1/in.)

Fig. 5-22(b) Curvature diagrams (PT specimen, FL2NC-e)

185
E W
M =-10.6 kN-m

- Mmid =1.13 kN-m - Mmid =1.13 kN-m

+ +
Moment Diagram Moment Diagram
M y =7.6 kN-m M y =7.6 kN-m
(yielding at one end) (yielding at one end)
M y =-12.8 kN-m M y=-12.8 kN-m

- Mmid =1.13 kN-m - Mmid =1.13 kN-m

Moment Diagram
+ Moment Diagram
+
M y =7.6 kN-m M y =7.6 kN-m
(yielding at both ends) (yielding at both ends)

Fig. 5-23 Slab moment diagrams within c + 3h at first yield of slab


reinforcement - RC specimen

E W
M =-14.2 kN-m
M =-7.6 kN-m
- -
+ +
M y =7.6 kN-m
M y=14.2 kN-m
Moment Diagram M y=-22.4 kN-m Moment Diagram
(yielding
kN-m at one end) (yielding at one end)
M y=-14.2

- -
+ +
My =7.6 kN-m
M y=14.2 kN-m
Moment Diagram Moment Diagram
(yielding at both ends) (yielding at both ends)

Fig. 5-24 Slab moment diagrams within column strip at first yield of slab
reinforcement - PT specimen

186
φ
φ

φ
r2 φ
r1

Fig. 5-25 Process used to compute slab rotation (RC specimen)

φ
φ
φ

φ
r2 φ
r1 φ

Fig. 5-26 Process used to compute slab rotation (PT specimen)

187
0.08 Conventional
Post-tensioned
with Stud-rails
RC- Int (Kang & Wallace)
RC- Ext (Kang & Wallace)

Rotation Capatity (Drift Ratio)


PT- Int (Kang & Wallace)
0.06 PT- Ext (Kang & Wallace)

Best-Fit Line :
Static Tests of
0.04 Conventional Slabs
with Stud-rails

Best-Fit Line :
Static Tests of
0.02 Post-tensioned Slabs

Best-Fit Line :
Static Tests of
Conventional Slabs
0
0 0.2 0.4 0.6 0.8 1
Gravity Shear Ratio (Vg/φVc)
Fig. 5-27 Drift ratio capacity with the existing database (See Table 5-2 and 5-3)

Fig. 5-28 Post-tensioning force distribution (at the completion of first post-tensioning)

188
Fig. 5-29 Post-tensioning force distribution (at the completion of third post-tensioning)

189
Fig. 6-1 Analytical model for displacement

Connection
( rigid plastic spring )
Column
( fiber element )

Slab ( elastic element ) Column strip


( rigid plastic spring )
(a) Frame

Munb Mu,unb Mslab


∑ M fy ,cs

Mn,unb γv = 1.0
My,unb Mfy,cs
Punching Yielding
Punching after Yielding within Column Strip,
prior to within c+3h Followed by
Yielding (item 2) Punching at θu - θ*
(item1) (item 3)

0 θ* θu θ 0 θ
(b) Rigid plastic spring for connection (c) Rigid plastic spring for column strip

Fig. 6-2 Inelastic model for rigid plastic springs

190
M P = PG + PE

P = PG
P = PG - PE

0.42 EIG

0.39 EIG

0.38 EIG
PG = axial from gravity
PE = axial from earthquake

φ
(a) Exterior column

P = PG

EIcr ≈ 0.45 EIG

PG = axial from gravity


PE = axial from earthquake

φ
(b) Interior column

Fig. 6-3 Crack coefficient for columns

191
Lateral Drift (%)
-0.4 -0.2 0 0.2 0.4
200
40

100
20

Base Shear (kips)


Base Shear (kN)

0 0

-20
-100 1.0 Ιcol, α=0.8, β=1
1.0 Ιcol, α=0.8, β=1/3
0.4 Ιcol, α=0.8, β=1/3
ATC (0.7 Ιcol, αβ=Κfp)
Measured (2nd Run)
-40
-200
-10 -5 0 5 10
Lateral Displacement (mm)

(a) RC specimen (Run 2-2) - Model A, B, C, D

Lateral Drift (%)


-0.4 -0.2 0 0.2 0.4
200
40

100
20
Base Shear (kips)
Base Shear (kN)

0 0

-20
-100 1.0 Ιcol, α=0.65, β=1
1.0 Ιcol, α=0.65, β=1/2
0.4 Ιcol, α=0.65, β=1/2
ATC (0.7 Ιcol, αβ=Κfp)
Measured (2nd Run)
-40
-200
-10 -5 0 5 10
Lateral Displacement (mm)

(b) PT specimen (Run 2) - Model A, B, C, D

Fig. 6-4 Top relative displacement vs base shear comparisons for elastic model

192
Top Displacement (mm)

Top Displacement (in)


20
Measured (Run 2)
Simulated 0.5
10

0 0

-10
-0.5
-20
0 10 20 30 40
Time (sec)
(a) RC (Run 2-2) - Model B
Top Displacement (mm)

Top Displacement (in)


20
Measured (Run 2)
Simulated 0.5
10

0 0

-10
-0.5
-20
0 10 20 30 40
Time (sec)
(b) RC (Run 2-2) - Model C
Top Displacement (mm)

Top Displacement (in)


20
Measured (Run 2)
Simulated 0.5
10

0 0

-10
-0.5
-20
0 10 20 30 40
Time (sec)
(c) PT (Run2) - Model B
Top Displacement (mm)

Top Displacement (in)

20
Measured (Run 2)
Simulated 0.5
10

0 0

-10
-0.5
-20
0 10 20 30 40
Time (sec)
(d) PT (Run 2) - Model C
Fig. 6-5 Top relative displacement comparisons for elastic model

193
Time (sec)
12 14 16
20
Measured (Run 2)
Simulated

0.5
10
Top Displacement (mm)

Top Displacement (in)


0 0

-10
-0.5

-20
12 14 16
Time (sec)

(a) RC (Run 2-2) - Model B ( 1.0 I col , α = 0.8, β = 13 )

Time (sec)
12 14 16 18
20
Measured (Run 2)
Simulated

0.5
10
Top Displacement (mm)

Top Displacement (in)

0 0

-10
-0.5

-20
12 14 16 18
Time (sec)

(b) RC (Run 2-2) - Model C ( 0.4 I col , α = 0.8, β = 13 )

Fig. 6-6 Top relative displacement comparisons at peak (RC)

194
Time (sec)
12 14 16
20
Measured (Run 2)
Simulated

0.5
10
Top Displacement (mm)

Top Displacement (in)


0 0

-10
-0.5

-20
12 14 16
Time (sec)

(a) PT (Run 2) - Model B ( 1.0 I col , α = 0.65, β = 12 )

Time (sec)
12 14 16
20
Measured (Run 2)
Simulated

0.5
10
Top Displacement (mm)

Top Displacement (in)


0 0

-10
-0.5

-20
12 14 16
Time (sec)

(b) PT (Run 2) - Model C ( 0.4 I col , α = 0.65, β = 12 )

Fig. 6-7 Top relative displacement comparisons at peak (PT)

195
Lateral Drift (%)
-0.4 0 0.4

50
200
α=0.8, β=1
α=0.8, β=1/3
Measured (2nd Run)

25
100

Base Shear (kips)


Base Shear (kN)

0 0

-100
-25

-200
-50

-10 0 10
Lateral Displacement (mm)

(a) RC specimen (Run 2-2) - Model E, F

Lateral Drift (%)


-0.4 0 0.4
200
40
α=0.65, β=1
α=0.65, β=1/2
Measured (2nd Run)
100
20
Base Shear (kips)
Base Shear (kN)

0 0

-20
-100

-40
-200

-10 0 10
Lateral Displacement (mm)

(b) PT specimen (Run 2) - Model E, F

Fig. 6-8 Top relative displacement vs base shear comparisons for fiber element model

196
Time (sec)
12 14 16
20
Measured (Run 2)
Simulated

0.5
10
Top Displacement (mm)

Top Displacement (in)


0 0

-10
-0.5

-20
12 14 16
Time (sec)
(a) RC (Run 2-2) - Model F (Fiber column, α = 0.8, β = 13 )

Time (sec)
12 14 16
20
Measured (Run 2)
Simulated

0.5
10
Top Displacement (mm)

0 0 Top Displacement (in)

-10
-0.5

-20
12 14 16
Time (sec)
(b) PT (Run 2) - Model F (Fiber column, α = 0.65, β = 12 )

Fig. 6-9 Top relative displacement comparisons at peak for fiber element model

197
Top Drift [%]
-4 -3 -2 -1 0 1 2 3 4
RC-RUN2-2
60 RC-RUN4
Push-Over - 0.5:1 ratio 1.5
(Opensees)
Push-Over - 2:1 ratio
40 (Opensees) 1

Base Shear [kips]

Base Shear [W]


20 0.5

0 0

-20 -0.5

-1
-40

-1.5
-60

-3.5 -3 -2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5 3 3.5

Top Relative Displacement [in.]

(a) RC specimen

Top Drift [%]


-5 -4 -3 -2 -1 0 1 2 3 4 5
PT-RUN2
PT-RUN4 1.5
60
PT-RUN5
Push-Over - 0.5:1 ratio
(Opensees)
40 1
Push-Over - 2:1 ratio
(Opensees)
Base Shear [kips]

Base Shear [W]

20 0.5

0 0

-20 -0.5

-40 -1

-60 -1.5

-4 -3 -2 -1 0 1 2 3 4

Top Relative Displacement [in.]


(b) PT specimen

Fig. 6-10 Push-over curves for top relative displacement vs base shear

198
Appendix A

(1) Stress - strain relations of #2 - 1 sample

199
(2) Stress - strain relations of #2 - 2 sample

200
(3) Stress - strain relations of #2 - 3 sample

201
(4) Stress - strain relations of #3 - 1 sample

202
(5) Stress - strain relations of #3 - 2 sample

203
(6) Stress - strain relations of #3 - 3 sample

204
(7) Stress - strain relations of #4 - 1 sample

205
(8) Stress - strain relations of #4 - 2 sample

206
(9) Stress - strain relations of #4 - 3 sample

207
(10) Stress - strain relations of Studrails - 1 sample

208
(11) Stress - strain relations of Studrails - 2 sample

209
(12) Stress - strain relations of Studrails - 3 sample

210
(13) Stress - strain relations of Strands - 1 sample

211
(14) Stress - strain relations of Strands - 2 sample

212

Você também pode gostar