Sunteți pe pagina 1din 262

All the material in this document is protected by copyright.

Any use of the work other than as authorized under copyright


law is prohibited.
University of Trento
University of Brescia
University of Padova
University of Trieste
University of Udine
University IUAV of Venezia






ANDREA BELLERI (Ph.D. Candidate)






DISPLACEMENT BASED DESIGN FOR PRECAST
CONCRETE STRUCTURES






Prof. Paolo Riva (Tutor)






SUMMARY

The objective of this research is to check the suitability of the Direct Displacement Based
Design (DDBD) procedure when applied to the seismic design of precast concrete
structures and how the procedure is affected by taking into account the influence of
foundation flexibility, beam to column and foundation to column connections.
Relationships are derived to relate the hysteretic parameters used to calibrate the
Equivalent Viscous Damping (EVD) equation (used in the DDBD procedure to estimate
the system damping) to the momentcurvature parameters model adopted to describe
the column flexural behavior. A set of experimental tests has been carried out to analyze
the behavior of some foundation to column connections typical of the precast industry
and to calibrate the parameters of the EVD equation associated to each connection type.
A new calibration procedure is proposed and applied to determine the parameters of a
modified formulation of the EVD equation, suitable for different hysteretic rules, and the
dependence of the latter on the type of ground motions (near field and far field) and on
the records spectral response shape is investigated.
The ground motion selection and scaling have been seen to play a significant role in the
design validation procedure, therefore a ground motions scaling procedure is proposed to
control and limit the results variability of non linear time history analyses and the results
dependence on the ground motion set chosen. This procedure could be suitable for the
EVD equation calibration other than for non linear time history analyses.
The study on precast concrete structures is extended to the use of rocking walls as an
alternative Lateral Force Resisting System (LFRS). The main issues related to this type of
system are outlined such as the base sliding, the definition of a moment rotation
relationship to use in the non linear analyses and the peculiarity of the equations of
motion. Rocking walls have been the lateral force resisting system of an extensive
experimental campaign on precast diaphragms involving shake table tests recently
concluded at the University of California at San Diego. The design and the experimental
results of these walls are presented and efforts have been made to explain their dynamic
behavior especially regarding the vertical and horizontal acceleration spikes associated to
the wall impacts once rocking is triggered.


SOMMARIO

Lobbiettivo principale della ricerca verificare lapplicabilit del metodo di progettazione
sismica chiamato Direct Displacement Based Design (DDBD) a strutture prefabbricate
tipiche del panorama costruttivo italiano. In particolare dinteresse definire come la
procedura viene modificata per tenere in considerazione gli effetti del terreno, delle
connessioni travi pilastro e delle connessioni tra pilastro e fondazione. Sono state
determinate le relazioni intercorrenti tra i parametri del modello isteretico usato per
calibrare lequazione dello smorzamento viscoso equivalente (EVD) nella procedura
DDBD e i parametri del legame momentocurvatura utilizzato per descrivere il
comportamento flessionale dei pilastri. Test sperimentali sono stati condotti per
analizzare il comportamento dal punto di vista sismico di alcune connessioni pilastro
fondazione prefabbricate e per calibrare i parametri dellequazione EVD ad esse
associati. Viene quindi proposta una procedura alternativa per la calibrazione dei
parametri di una nuova formulazione dellequazione EVD, adatta a vari modelli isteretici,
valutando la dipendenza di questultima dalle caratteristiche degli accelerogrammi
adottati, se di tipo near field o far field, e della relativa forma dello spettro di risposta.
La scelta e lo scaling degli accelerogrammi giocano un ruolo significativo in fase di
validazione non lineare della progettazione. In questa sede proposta una procedura di
scaling atta a controllare e limitare la variabilit dei risultati delle analisi non lineari dovuta
alla scelta degli accelerogrammi. In particolare tale procedura pu essere applicata nella
calibrazione dei parametri dellequazione EVD.
Dopo avere esaminato sistemi sismo resistenti classici utilizzati nella prefabbricazione,
vale a dire costituiti da pilastri isostatci considerando o meno leffetto delle connessioni
trave-pilastro, la ricerca continua con lo studio di sistemi sismo resistenti innovativi,
quali lo sono i rocking walls, delineandone le caratteristiche e gli aspetti principali. Sono
presentati la progettazione e i risultati sperimentali dei rocking walls utilizzati come
sistema sismo resistente in una campagna di prove su tavola vibrante recentemente
conclusa allUniversit della California San Diego. Tali risultati permettono considerazioni
significative riguardanti il comportamento di questi sistemi quando sottoposti ad
eccitazione dinamica, in particolare i picchi nellaccelerazione verticale e orizzontale a cui
sono soggetti.


ACKNOWLEDGEMENTS


I gratefully thank all the people who sustained me in these last three years and the
people who helped directly or indirectly to extend my knowledge.
My thanks and appreciation to my advisor prof. Paolo Riva who gave me the opportunity
to take this challenge and helped me handling it.
I thank all the friends, colleagues, technicians and professors who made pleasant and
fruitful my year and a half permanence at the University of California at San Diego
(UCSD). At this regard special thanks to Matthew Schoettler for having been a friend and
a guide and to prof. Jos Restrepo who supervised the shake table experimental tests
and the last part of this research.






alla mia famiglia



INDEX


1. INTRODUCTION ............................................................................................................. 1
1.1 Displacement Based Design methodologies ........................................................... 4
1.2 Research plan description ....................................................................................... 9
2. GROUND MOTIONS AND CASE STUDIES ..................................................................... 13
2.1 Ground motions definition .................................................................................... 13
2.2 Case studies definition .......................................................................................... 14
3. FBD AND DDBD PROCEDURES ..................................................................................... 19
3.1 FBD Procedure ....................................................................................................... 19
3.2 DDBD Procedure .................................................................................................... 20
3.3 Considerations on equivalent viscous damping .................................................... 23
3.3.1 Relationship between displacement (

) and curvature (

) ductility .............. 27
3.3.2 Relationship between r and r ............................................................................ 28
3.3.3 Relationship between and ........................................................................ 29
3.3.4 Relationship between and .......................................................................... 30
3.4 Displacement response spectrum dependence from damping ............................ 31
4. PRECAST STRUCTURES CONSIDERATIONS ................................................................... 33
4.1 Definition of the q-factor for precast concrete buildings ..................................... 33
4.2 Precast concrete structures compared to other structures .................................. 35
5. PHASE 1: FBD AND DDBD COMPARISON ..................................................................... 39
5.1 FBD-DDBD comparison: Procedure 1 .................................................................... 40
5.2 FBD-DDBD comparison: Procedure 2 .................................................................... 51
5.3 Considerations about the inelastic displacements ............................................... 56
5.4 DDBD for 2.5% drift ............................................................................................... 59
5.5 Concluding remarks ............................................................................................... 61


6. PHASE 2: DDBD AND SOIL STRUCTURE INTERACTION ................................................ 63
6.1 First approach: elastic foundation ........................................................................ 64
6.2 Second approach: inelastic foundation ................................................................. 67
6.3 Considering foundation inertia ............................................................................. 73
6.4 Concluding remarks ............................................................................................... 77
7. PHASE 3: DDBD AND BEAM TO COLUMN CONNECTION ............................................ 79
7.1 Analytical study ..................................................................................................... 79
7.2 Procedure application to the case studies ............................................................ 90
7.3 Concluding remarks ............................................................................................... 98
8. PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION ................................ 99
8.1 Experimental tests ................................................................................................. 99
8.2 DDBD application ................................................................................................ 116
8.3 Yield curvature equation ..................................................................................... 118
8.4 Equivalent viscous damping equation re-calibration and results ....................... 125
8.5 Concluding remarks ............................................................................................. 130
9. HYSTERTIC DAMPING EQUATION CALIBRATION ....................................................... 131
9.1 Hysteretic damping calibration results ............................................................... 132
9.2 Hysteretic model parameters influence on the damping value .......................... 142
9.3 Concluding remarks ............................................................................................. 152
10. GROUND MOTION SCALING .................................................................................... 153
10.1 Record selection and scaling ............................................................................. 153
10.2 Constant Variance Spectrum Matching procedure ........................................... 155
10.3 CVSM application .............................................................................................. 158
10.4 Concluding remarks ........................................................................................... 170
11. ROCKING WALLS IN PRECAST CONCRETE STRUCTURES .......................................... 171
11.1 Rocking walls: an introduction .......................................................................... 171
11.2 Rocking walls experimental tests in the literature ............................................ 172
11.3 Rocking wall base sliding ................................................................................... 175
11.4 Equations of motion .......................................................................................... 179
11.5 Design recommendations ................................................................................. 184
11.6 Non linear time history analyses ....................................................................... 185


12. SHAKE TABLE TESTS INVOLVING ROCKING WALLS ................................................. 191
12.1 Hybrid wall general considerations and details ................................................ 193
12.2 Hybrid wall design ............................................................................................. 200
12.3 Test sequence .................................................................................................... 204
12.4 Instrumentation layout ..................................................................................... 207
12.5 Tests results ....................................................................................................... 210
12.6 Concluding remarks ........................................................................................... 232
13. CONCLUSION AND FUTURE DEVELOPMENTS ......................................................... 233

BIBLIOGRAPHY ................................................................................................................ 237

APPENDIX A: SECTION DATA FOR NONLINEAR ANALYSES .............................................. 245


LIST OF SYMBOLS



A
d
dissipation bar area
A
PT
post tensioning strand area
B column cross section size
b
e
confined concrete region thickness
B
f
foundation dimension
c
cover
concrete cover
c
d
viscous damping coefficient
c
NA
neutral axis depth
d
e
elastic displacement
d
i
inelastic displacement
E
c
concrete Young modulus
E
dissipated
energy dissipated in one cycle
E
elastic
elastic energy at maximum response
E
s
steel Young modulus
f
cc
confined concrete strength
f
ck
concrete cylindrical strength
F
d
damping force
F
e
elastic force
F
i
i-floor design force
F
l
maximum confining lateral stress
f
l
minimum confining lateral stress
F
p0
initial prestress
F
PT
post tensioning force
F
u
ultimate lateral force


F
y
yield lateral force
f
yk
steel yield stress
g acceleration of gravity
G soil shear modulus
G
red
reduced soil shear modulus
H structure height
H
eff
structural effective height
I
eff
effective modulus of inertia
I
gross
gross modulus of inertia
k structural stiffness
k
eff
effective stiffness
k
i
initial stiffness force-displacement relationship
k
i
initial stiffness moment-curvature relationship
K
s
superstructure stiffness
k
u
unloading stiffness force-displacement relationship
k
u
unloading stiffness moment-curvature relationship
K
x
foundation horizontal stiffness
k
y
yield stiffness force-displacement relationship
K
z
foundation vertical stiffness
K

foundation rotational stiffness


L
p
plastic hinge length
l
unb_
dissipation bar unbonded length
l
unb_PT
tendon unbonded length
L
w
wall depth
m seismic mass
m
eff
effective seismic mass
M
u
design moment
M
y
yield moment


N axial load
P gravity load
q force reduction factor
r post-yield stiffness ratio force-displacement relationship
r post-yield stiffness ratio moment-curvature relationship
S
a
spectral acceleration
S
D
spectral displacement
T
0
structural period at secant stiffness at yield
T
eff
effective period
V
b
base shear
v
s
shear wave velocity


GREEK SYMBOLS



Takeda model parameter force-displacement relationship
Takeda model parameter moment-curvature relationship
Takeda model parameter force-displacement relationship
Takeda model parameter moment-curvature relationship

d
target displacement

f
displacement due to foundation rotation

res
residual displacement

s
structural displacement

u
ultimate displacement

y
inelastic displacement

cu
maximum concrete compressive strain

y
steel yield strain

p
plastic curvature

res
residual curvature

u
ultimate curvature

y
yield curvature
spectrum damping dependence

displacement ductility

curvature ductility
axial load ratio

soil
soil Poisson modulus
second to first order moment ratio

f
foundation rotation

0
yield to gross stiffness ratio

l
longitudinal steel ratio

soil
soil density
angular frequency

f
foundation angular frequency

s
structure angular frequency

init
el
initial stiffness elastic damping

tang
el
tangent stiffness elastic damping

eq
equivalent viscous damping

f
foundation equivalent viscous damping

hyst
hysteretic damping

s
structural equivalent viscous damping


1
1. INTRODUCTION
















Several efforts have been made in the last decades to address the importance of
changing the focus of current seismic design codes from merely preventing collapse in
major earthquakes and controlling the damage in minor earthquakes to a more general
design philosophy which takes into account multiple performance objectives based on
quantifiable performance criteria; this design philosophy is referred to as Performance
Based Design (PBD). The Olive View Hospital in Sylmar (CA - USA) represents one of
the most significant examples of the need of PBD rather than Force Based Design (FBD)
approach adopted by current codes. The hospital was destroyed by the 1971 San
Fernando earthquake and it was completely rebuilt in 1976 to withstand increased levels
of seismic forces according to a life safety criterion. The lateral force resisting system
adopted is a mix design of concrete and steel shear walls which resulted in a very strong
and stiff structure. During the 1994 Northridge earthquake the sensors in the building
indicated a peak ground level acceleration of 0.91 g sensibly beyond the design
acceleration of 0.52 g at which the building would not be damaged badly. The roof peak
acceleration was recorded to 2.31 g. From a structural point of view the hospital did not
sustain damage, but it had to be evacuated because of broken water pipes and other
secondary damage with sensible economic losses. This example underlines the





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
2
inadequacy of design practice targeted only to life safety and collapse prevention criteria
and the important role of nonstructural components in the functionality of a building after
an earthquake: this requires the development and application of PBD methodologies.
The performance objectives are statements that relate an acceptable performance level
in a structure to an earthquake design level. Performance targets which can be specified
limits on response parameters (like stresses, strains, displacements and accelerations
among others) correspond to each performance level. As mentioned before, particular
care has to be placed to non structural performance levels as well because the way the
non structural components (like partitions, ceilings, elevators and electrical, plumbing,
mechanical, and fire protection systems) behave during an earthquake will affect the
building operability and occupancy following an earthquake.
The SEAOC (1995) made an effort to relate the performance levels to the expected
damage in the overall building (i.e. both structural and non structural elements) as
summarized in Table 1.1.
Table 1.1. Performance Levels and Damage States
Performance
Levels
Damage States
Fully
Operational
No damage. Continuous service: facilities operate right after
earthquake.
Operational The structure is safe for occupancy immediately after earthquake.
Repair is required to restore some essential services.
The structure retains a significant portion of its original stiffness and
most of its strength
Life Safe Life safety is attained and the structure remains stable although
damaged. Substantial damage has occurred to the structure, and it
may have lost a significant amount of its original stiffness.
Significant margin remains before collapse would occur.
Near Collapse Severe damage. Non structural elements may fall. If laterally
deformed beyond this point, the structure can experience instability
and collapse
Collapse Portions of primary structural system collapse. Or as extreme the
whole structure collapses.






INTRODUCTION
3
A performance objective is a coupling of expected performance levels with levels of
seismic hazard, which is represented, at a given site, as a set of earthquake ground
motions with specified probabilities of occurrence. SEAOC (1995) relates four levels of
seismic hazard to three sets of performance objectives, which are associated to three
types of facilities: Basic Facilities, Essential Facilities and Safety Critical Facilities
(Table 1.2).
Table 1.2. Performance Levels and Damage States
Objectives Earthquake Performance Level
Fully
Operational
Operational Life Safe Near Collapse
E
a
r
t
h
q
u
a
k
e

D
e
s
i
g
n

L
e
v
e
l

Frequent
(50% in 30 years)
Basic
Facilities
Unacceptable performance
Occasional
(50% in 50 years)
Essential
Facilities
Basic
Facilities
(for new constructions)
Rare
(5% in 50 years)
Safety Critical
Facilities
Essential
Facilities
Basic
Facilities

Very Rare
(2% in 50 years)
Safety Critical
Facilities
Essential
Facilities
Basic
Facilities

Therefore the design method chosen has to start from the definition of the performance
objective associated to the appropriate earthquake design level and performance level.
The performance levels should be defined by parameters which allow a quantitative
identification of the structural performance and the design methodology should deal
directly with these parameters. A seismic code which addresses the seismic design in
terms of equivalent seismic forces as done in the past is not suitable for the procedure
just described, because the main design parameter induced by the code scheme is the
structural strength which is not directly correlated to damage.
The performance targets could be a level of stress not to be exceeded, a load, a
displacement or a limit state. The target displacement was first applied as the response
parameter (Displacement Based Design) because the structural response in terms of
displacement can be related to strain based limit states, which give better indicators of
damage than stresses. Based on these considerations Performance Based Design and
Displacement Based Design have been used interchangeably. This assumption is an
oversimplification since the level of damage is influenced by several other parameters like





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
4
for example the number of cycles and the duration of the earthquake and the acceleration
levels which affect the behavior of secondary systems. Displacement Based Design
should be thought as a subset of Performance Based Design.

1.1 Displacement Based Design methodologies
A quick review of the main Displacement Based Design (DDBD) methodologies available
in the literature is presented in this paragraph. According to FIB bulletin 25 (2003) the
design procedures involving DBD can be organized by different criteria that will be used
here as reference in the procedures presented.
The first distinction is the type of analysis used in the design process:
1. Initial Stiffness Response Spectra (ISRS): the procedure utilizes elastic stiffness
coupled with approximations between elastic and inelastic response.
2. Secant Stiffness Response Spectra (SSRS): the procedure adopts the secant
stiffness to the maximum response and the concept of equivalent viscous
damping, which will be addressed in Chapter 3.
3. Time History Analysis (THA): the procedure uses linear or non-linear time history
methods to solve the equations of motion by direct integration for a given
earthquake in order to evaluate the system maximum response.
The role of the displacement in the design process is the second distinction taken into
account:
1. Displacement Calculations Based (DCB): the procedure involves the calculation
of the maximum displacement for an already designed structure. Detailing is
made to lead to a displacement capacity greater than the demand, but no attempt
is made to alter the structural system in order to change the displacement
demand.
2. Iterative Displacement Specification Based (IDSB): the procedure involves the
maximum displacement calculation for a designed structure as before, but
iterative changes are made on the structural system in order to limit the maximum
displacement to a specified value.
3. Direct Displacement Specification Based (DDSB): the procedure involves a
specified target displacement as a starting point. The structure design follows
directly leading, as end results, to the structural strength and stiffness necessary
to reach the target displacement under the specified earthquake level.






INTRODUCTION
5
Panagiotakos & Fardis (1999) proposed a displacement-based seismic design of
multistory reinforced concrete buildings integrated into the overall structural design
process including the effects of gravity loads. The procedure is summarized in Table 1.3.
Table 1.3. Panagiotakos & Fardis procedure
Type of analysis used in the design process ISRS
Role of displacement in the design process DCB
Description:
(1) The first step of the procedure is the elastic analysis for non-seismic actions and
serviceability earthquake with an elastic spectrum and adopting un-cracked sections.
(2) Then a force-based proportioning of the longitudinal reinforcement in hinge
locations is carried out and the capacity design rule is applied throughout the structure.
(3) An elastic analysis for life-safety earthquake is carried out with a 5% damped
spectrum and using the secant to yield members stiffness. (4) The upper-
characteristics of chord-rotation demands are evaluated with provided amplification
factors obtained from extensive time history analyses. (5) The chord rotation demand
is verified and the longitudinal and transverse re-bars are modified if necessary. (6)
Finally capacity design is applied.

Freeman (1998) proposed to compare response spectra for different levels of damping
with the capacity spectrum obtained from dynamic considerations on pushover analysis
results of a multi degree of freedom (MDOF) system (Table 1.4).
Table 1.4. Freeman procedure
Type of analysis used in the design process SSRS
Role of displacement in the design process DCB
Description:
(1) For a given MDOF system determine through a pushover analysis the system
capacity curve in terms of roof displacement versus base shear. (2) Use the dynamic
characteristics of the structure (such as period of vibrations, mode shapes and modal
participation factors) to convert (3) the MDOF capacity curve to a capacity spectrum in
terms of Spectral acceleration versus Spectral displacement. (4) Calculate the
response spectra for various levels of damping. (5) Use ductility-damping relation to
identify different damping levels along the capacity spectrum curve. (6) The
intersection of the capacity spectrum with the response spectrum with the appropriate
level of damping determine the seismic structural demand.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
6
Fajfar (2000) modified the capacity spectrum method proposed by Freeman adopting
inelastic instead of elastic response spectra in the procedure (Table 1.5).
Table 1.5. Fajfar procedure
Type of analysis used in the design process ISRS
Role of displacement in the design process DCB
Description:
(1) For a given MDOF system determine through a pushover analysis the system
capacity curve in terms of roof displacement versus base shear. (2) Divide the force
and displacement obtained by the modal participation factor of the first mode of
vibration. This will determine the force and displacement of the equivalent single
degree of freedom (SDOF) system. (3) Calculate the inelastic response spectra
associated to different ductility values. (4) Use the capacity spectrum and the response
spectra to determine the displacement demand of the SDOF system. (5) Convert the
SDOF displacement demand into MDOF maximum top displacement.

Aschheim & Black (2000) procedure differs from the previous ones because it involves
the use of yield point spectra representing the yield points of oscillators with constant
displacement ductility (Table 1.6)
Table 1.6. Aschheim & Black procedure
Type of analysis used in the design process ISRS
Role of displacement in the design process DDSB
Description:
(1) Develop yield point spectra for various ductility levels. (2) Determine target
displacement that satisfies limits for desired risk event. (3) Identify in the yield point
spectra the admissible design region in terms of system displacement and ductility.
(4) Choose ductility limit for the desired performance level. (5) Determine the structural
yield displacement and the corresponding yield strength. (6) Distribute the lateral force
according to conventional methods. (7) Design and detail the structure.






INTRODUCTION
7
Browning (2001) procedure allows to design regular reinforced concrete frame structures
to reach a predefined average drift limit (Table 1.7).
Table 1.7. Browning procedure
Type of analysis used in the design process ISRS
Role of displacement in the design process IDSB
Description:
(1) The first step is the evaluation of the maximum target period, whose exceedance
will result in the drift exceeding a specified value, using displacement response
spectra. (2) Proportion the members based on gravity load requirements. (3) Adjust
the member size until the structural period is less than the target period. (4) Evaluate
the base shear from structural period and compare it to an acceptable minimum. (5)
Ensure an appropriate hierarchy of strength. (6) Provide structural details compatible
with the maximum tolerable drift.

The displacement based procedure proposed by Priestley (1997) adopts a substitute
structure approach to characterize the structure by a single degree of freedom system
with stiffness the secant structural stiffness at maximum displacement and with a level of
equivalent viscous damping appropriate to take into account the hysteretic energy
absorbed during the inelastic response.
Table 1.8. Priestley procedure
Type of analysis used in the design process SSRS
Role of displacement in the design process DDBS
Description:
(1) Estimate the yield deformation of the system (first inelastic mode of vibration).
(2) Determine the SDOF substitute structure effective height and effective mass from
the system deformed shape at yield. (3) Determine the system equivalent viscous
damping to represent the elastic and hysteretic damping of the system. (4) Get the
substitute structure effective period from the displacement spectrum reduced
according to the equivalent viscous damping. (5) Determine the substitute structure
effective stiffness and the system base shear. (6) Distribute the base shear as design
forces along the structure proportionally to the inelastic displacements and masses. (7)
Design the structure member according to the design forces and apply capacity
design.






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
8
Chopra & Goel (2001) proposed a modified version of the method of Priestley (1997)
which adopts inelastic spectra (Table 1.9).
Table 1.9. Chopra & Goel procedure
Type of analysis used in the design process ISRS
Role of displacement in the design process DDBS
Description:
(1) Estimate the yield deformation of the system.
(2) Determine the design displacement and ductility factor from acceptable plastic
rotation considerations.
(3) Enter inelastic constant ductility displacement spectra with the system
displacement and ductility to get system initial period.
(4) Determine the system initial stiffness and the required yield strength.
(5) Estimate member sizes and detailing to provide the system strength required.

The procedure proposed by Restrepo (2007) introduces in the design additional factors to
take into account ground motion variability and the inelastic versus elastic displacement
demand variability (Table 1.10).
Table 1.10. Restrepo procedure
Type of analysis used in the design process ISRS
Role of displacement in the design process DDBS
Description:
(1) Select an appropriate mechanism of inelastic deformation. (2) Select the level of
detailing in the plastic hinge regions. (3) Calculate the reference yield displacement.
(4) Calculate the theoretical ultimate lateral displacement. (5) Determine the
displacement ductility and the C
QR
coefficient (which accounts for ground motion
variability and inelastic versus elastic displacement demand variability). (6) Scale the
elastic displacement spectrum by C
QR
and determine the period correspondent to the
system ultimate displacement. (7) Determine the base shear and distribute it along the
height of the building. (8) Complete the design and apply capacity design.






INTRODUCTION
9
The procedure proposed by Kappos & Manafpour (2001) is the only one presented here
involving time history analyses as part of the procedure (Table 1.11)
Table 1.11. Kappos & Manafpour procedure.
Type of analysis used in the design process THA
Role of displacement in the design process DCB
Description:
(1) Apply force based design to obtain a basic strength level for serviceability
earthquake combined with gravity loads. (2) Detail of the beams flexural reinforcement.
(3) Use non linear time history analyses to check maximum drifts and plastic rotations
in beam critical regions associated to an earthquake with probability of exceedance
50% in 50 years (beams modeled as yielding elements while columns as elastic ones).
(4) Scale the ground motions to an event with correspondent probability of exceedance
10% in 50 years. (5) The time history analyses provide the critical moment and axial
load combination at each column critical section. (6) Detail the column longitudinal
reinforcement. (7) Design and detailing of all members for shear. (8) Detail all
members for confinement, anchorages and lap splices.

In this research the procedure of Priestley (1997), usually referred as Direct Displacement
Based Design (DDBD), has been adopted mainly for two reasons: the first one is that it
involves a specified target displacement as a starting point, which is seen as a suitable
design procedure; the second reason is that several efforts have been recently made to
implement the aforementioned procedure in the Italian Seismic Design Code.

1.2 Research plan description
The main objective of this research is to check the suitability of the DDBD procedure
when applied to the seismic design of precast concrete structures and how the procedure
is affected by taking into account the influence of foundation flexibility, the different types
of foundation to column connections and beam to column connections typical of precast
buildings. The typical structural layout of Italian warehouses and commercial malls
consists of concrete cantilever columns, connected by simply supported precast and
prestressed beams, supporting prestressed concrete roof elements; the columns are
inserted and grouted in place in isolated precast cup-footings; reducing the construction
time, this solution is extremely cost effective.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
10
The case studies examined and the earthquake records adopted in the non linear time
history analyses to validate the DDBD procedure are shown in Chapter 2.
The DDBD and the Force Based Design (FBD) procedures are presented in Chapter 3;
considerations are made on the equivalent viscous damping equations available in the
literature and relationships are derived to relate those equations, which have been
derived from force-displacement analyses of single degree of freedom systems, to the
moment curvature relationship actually adopted in the column flexural description for
the non linear time history analyses necessary to validate the procedure.
In Chapter 4, specific considerations on the application of DDBD procedure to the precast
concrete structures considered are made. Compared to traditional reinforced concrete
structures the typology under exam presents lower displacement ductility demand due to
the higher interstory height; this suggests to check the implications of the equivalent
viscous damping equations, which have been derived for larger displacement ductility
values, when applied to these structures.
The suitability of the DDBD procedure when applied to the seismic design of precast
concrete structures is evaluated through four phases whose schematic representation is
shown in Figure 1.1.
In Phase 1, Chapter 5, the equations developed in Chapter 3 are used for the comparison
between FBD and DDBD procedures to outline advantages and drawbacks and the
conditions under which the two procedures give compatible results. Although the
comparison in not straightforward, due to the different inelastic displacement
computations, two possible ways of doing it are proposed. The moment curvature
relationship adopted to describe the columns flexural behavior is the same used to
calibrate the equivalent viscous damping equations available in the literature. At the end
of the chapter the DDBD is applied to precast structures with a different moment-
curvature relationship showing the need of a more rigorous calibration of the equivalent
viscous damping equations.
In Phase 2, Chapter 6, the influence of Soil Structure Interaction is taken into account in
the DDBD procedure adding the foundation flexibility and damping limited to the
foundation rocking motion. Two possible ways of doing it are taken into account,
extending results available in the literature, which consider a single degree of freedom
substitute structure obtained by static condensation without considering the foundation
inertia. Both analytical and non linear analyses are carried out to check the procedure
suitability when foundation inertia is considered.






INTRODUCTION
11
Phase 1 Phase 2
keff , eq
H
=
H
e
f
fm=meff
d

d
Kx
Kz
K
H
=
H
e
f
f
keff , eq
m=meff

Phase 3 Phase 4
H
H
e
f
f
keff, eq
d
m
meff

keff , eq
H
=
H
e
f
fm=meff
d


Figure 1.1 Schematic approach of the DDBD procedure validation.

In Phase 3, Chapter 7, the influence of the top connection in the DDBD procedure is
evaluated. The implementation of this aspect in the DDBD led to consider a substitute
structure with an effective height corresponding to the point of counter flexure and with as
effective mass the whole system mass. The analytical procedure developed is then
applied to the case studies selected.
In Phase 4, Chapter 8, the DDBD procedure is applied to systems whose hysteretic
behavior is different from the ones used in the equivalent viscous damping equation
calibration especially considering some of the foundation to column connections adopted
in the precast industry. Experimental tests are carried out to compare different types of
connections from a seismic and retrofitting point of view and to determine their hysteretic





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
12
parameters to use in the equivalent viscous damping equation calibration. A new and
faster calibration algorithm is proposed as alternative to the one available in the literature.
The calibration procedure is carried out for the hysteretic relationships associated to the
experimental tests and the results applied to the DDBD procedure. A new equation to
relate the yield curvature to the column cross section effective depth and the axial load
ratio is proposed; this formulation overcomes the drawback of the equation available in
the literature especially when applied to some foundation to column connections typical of
the precast industry. The equivalent viscous damping equation procedure is extended in
Chapter 9 to other hysteretic rules. A new equation is proposed and the influence of the
hysteretic parameters, the ground motion types and the displacement response spectrum
shape is evaluated.
In Chapter 10 a ground motions scaling procedure is proposed to control and limit the
coefficient of variation (defined as the standard deviation divided by the mean value) of
the acceleration response spectrum of the records set chosen, in order to limit the results
variability of non linear time history analyses and the results dependence on the ground
motion set chosen. This procedure seems suitable for the equivalent viscous damping
equation calibration procedure other than for non linear time history analyses.
After the application of DDBD procedure to precast structures with classical lateral force
resisting systems (i.e. fixed end columns with or without the contribution of the top column
to beam connection), Chapter 11 exploits the use of rocking walls as an alternative
resisting system to use in precast structures. This system has self centering properties
(given by post tensioning unbonded tendons) and accommodates the seismic lateral
displacement demand with a base rotation which leads to only one concentrated opening
of the foundation to wall joint compared to the crack spreading and damage typical of the
plastic region of classical reinforced concrete walls. This chapter deals with the problem
of the base sliding typical of these walls, with the definition of a moment rotation
relationship to use in the non linear analyses and with the revisiting and extension of the
equations of motion especially to determine the rocking period of the system whose
relation to the design procedure can be exploited as an extension of this research.
Rocking walls have been the lateral force resisting system of an extensive experimental
campaign on precast diaphragms recently concluded at the University of California, at
San Diego, and involving shake table tests. The design and the experimental results of
these walls are shown in Chapter 12 where efforts have been made to explain their
dynamic behavior in particular regarding the vertical and horizontal acceleration spikes
associated to the wall impacts once rocking is triggered.

13
2. GROUND MOTIONS AND CASE STUDIES


This chapter presents the ground motions and the case studies adopted to check the
suitability of the Direct Displacement Based Design (DDBD) procedure when applied to
the seismic design of precast concrete structures.

2.1 Ground motions definition
The elastic spectrum used in the design procedure, according to Eurocode 8-1:2004, is
the type 1 spectrum for a soil type C with a peak ground acceleration of 0.5 g. To validate
the DDBD procedure by means of non linear time history analyses, both natural and
artificial ground motions (Table 2.1) have been adopted and scaled, multiplying the
acceleration record by a scale factor, in order to match the Eurocode 8 design spectrum.
Table 2.1 Time history definition
Name Origin/Earthquake Duration (s) t (s) Scale factor
TH1 Duzce 25.89 0.01 1.2
TH2 Kalamata 29.995 0.005 3.1
TH3 Kocaeli 1 70.38 0.02 2.1
TH4 Northridge Baldwin 60.00 0.02 4.5
TH5 Hella 60.000 0.005 2.0
TH6 SIMQKE Aritif4 19.99 0.01
TH7 SIMQKE Aritif6 19.99 0.01

As Eurocode 8 states (3.2.3.1.2.4b): in the range of periods between 0.2T
1
and 2T
1
,
where T
1
is the fundamental period of the structure in the direction where the ground
motion will be applied, no value of the mean 5% damping elastic spectrum, calculated
from all time histories, should be less than 90% of the corresponding value of the 5%
damping elastic response spectrum. In the case under consideration it is possible to note
how the previous requirement is satisfied (Figure 2.1), so the ground motion records
adopted seem suitable. It is important to note that the constant displacement predicted by
the Eurocode 8 equation after the corner period of 2 s is not respected by the mean
displacement spectrum of the records used; therefore particular care should be taken in





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
14
the validation of the procedure in the case the substitute structure effective period is
greater than the corner period of Eurocode 8 spectrum.

Acceleration spectra comparison
0.0
0.5
1.0
1.5
2.0
2.5
0.0 1.0 2.0 3.0 4.0
Period (s)
A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
EC8
GM mean

Displacement spectra comparison
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
0.0 1.0 2.0 3.0 4.0
Period (s)
D
i
s
p
l
a
c
e
m
e
n
t

(
m
)
EC8
GM mean

Figure 2.1 Acceleration and Displacement response spectra comparison

2.2 Case studies definition
The case studies chosen are three existing buildings whose structural layout is typical of
Italian precast structures:
1. One story precast concrete building with double tee roof elements (Figure 2.2).
2. One story precast concrete building with omega roof elements (Figure 2.3).
3. One story building with precast concrete columns connected at the top by wood
beams (Figure 2.4 and Figure 2.5).





GROUND MOTIONS AND CASE STUDIES
15

Plan View
A
A
B
B
1
7
5
0
1
7
5
0
8
7
8
5
1
7
5
0
1090
7630
1090 1090 1090 1090 1090 1090
1
7
5
0
1
7
5
0

Section AA Section BB
C
L
7
5
0
8,40 m
1090 1090 1090
0,00 m

C
L
7
5
0
9
5
7
5
0
1750 1750
0,00 m

Double Tee Beam Details

18
60
250
51
5
6
0
60
51
130
112 18

Figure 2.2 - Case Study 1: one story building with double tee roof beams.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
16

Plan View
9
1
6
9
1
7
2000 2000
9
1
7

Section AA
5
1
0
7
0
1
4
5
1
2
5

Section BB Omega Beam Detail
7
3
5
6
3
5
1
0
0



Figure 2.3 - Case Study 2: one story building with omega roof beams.





GROUND MOTIONS AND CASE STUDIES
17

Plan View

CL

Figure 2.4 - Case Study 3: one story building with timber beam precast column connections.
A A





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
18

Section AA



Timber Beam concrete column connection detail

Two dowels connection Four dowels connection
Figure 2.5 - Case Study 3: timber beam precast column connection details.


19
3. FBD AND DDBD PROCEDURES


In this chapter the Direct Displacement Based Design (DDBD) and the Force Based
Design (FBD) procedures are presented and considerations are made in order to allow
the procedures comparison in Chapter 5. The issue of the equivalent viscous damping
(used in DDBD) is taken into account: relationships are derived to relate equations
available in the literature to the moment curvature relationship actually adopted in the
column flexural description, which will be used in the non linear time history analyses
necessary to validate the design. The dependence of the elastic response spectra of the
ground motion selected from the damping value is also checked.

3.1 FBD Procedure
For sake of simplicity the FBD procedure shown here corresponds to the case of a single
degree of freedom system, which will be adopted in the FBD and DDBD procedures
comparison. The FBD procedure is therefore:
1. Define a force reduction factor (q) for the structure.
2. Define an effective modulus of inertia I
eff
as a percentage of the gross modulus
I
gross
.
3. Define the stiffness of the system and determine the system period:

2
2
m
T
k

= = (3.1)
4. Determine the spectral acceleration corresponding to the structural period from
the design spectrum, ( )
a
, S T q .
5. Determine the base shear and the base moment as:

a b
V S m g = ;
u b
M V H = (3.2), (3.3)
6. Find the corresponding top displacement. The inelastic displacement, according
to Eurocode 8 4.3.4, is evaluated as:
i d e
d q d = , where q
d
is taken equal to q:
( ) /
b
q V k = (3.4)





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
20
7. Evaluate the second order effects computing the value, defined as the ratio
between the second order moment and the moment from analysis.

i
b
P d
V h

(3.5)
- less than 0.1: second order effects negligible.
- between 0.1 and 0.2: the base moment becomes ( ) ( ) / 1
u b
M V H = .
- between 0.2 and 0.3: second order effects to be taken into account.
- greater than 0.3: change structural dimensions.

3.2 DDBD Procedure
1. Definition of a single degree of freedom substructure system (Figure 3.1).


With:

d
target displacement
H
eff
effective height
m
eff
effective mass

eq
equivalent viscous damping
k
eff
effective stiffness
Figure 3.1 Substitute structure for DDBD procedure.

2. Determine the target displacement (
d
).
The target displacement depends from both the structural deformed shape and
the limit state under consideration, whose critical value can be associated either
to structural components (related to material strains) either to non-structural
components (related to interstory drift).
A linear distribution of the yield curvature from the column base to the top
(considered as the roof mass centroid) has been considered in Phase 1 of this
research (Chapter 5); this represents only a first approximation, without
considering the moment at top of the column due to beam connection (as it will
H
eff

meff

d

k
eff
,
eq





FBD AND DDBD PROCEDURES
21
be analyzed in Phase 4 Chapter 7).
A design based on the damage limitation requirement has been adopted and
the critical value of the target displacement is the one associated to an interstory
drift of 2.5 %.
To determine the yield displacement (
y
) the following equations have been used
(Priestley 2003):
2.1
y
y
B

= ,
2
2
2 3 3
y y
y
H H
H

= =
(3.6), (3.7)
Where:

y
is the yield curvature
B is the column cross section size

y
is the steel yield strain
3. Determine the effective height (H
eff
).
The effective height is the point where the system ductility is evaluated; it is
defined as:

1
1
n
i i i
i
eff n
i i
i
m H
H
m
=
=

(3.8)
4. Determine the effective mass (m
eff
).
The effective mass represents the mass participating in the first inelastic mode of
vibration and it is obtained considering the design displacement profile
i
of the
masses m
i
at each floor:

1
d
n
i i
i
eff
m
m
=

(3.9)
5. Determine the equivalent viscous damping (
eq
).
The DDBD adopts an equivalent viscous damping approach to represent the
elastic and the hysteretic damping of the system
eq
=5+
hyst
%; the first term, the
elastic viscous damping, takes into account material viscous damping, radiation
damping due to the foundation system and damping due to the non linear
behavior of the connections.
The second term, the hysteretic damping, depends on the hysteretic relationship





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
22
of the structural elements and takes into account in somehow the capacity of the
system to dissipate energy.
6. From the equivalent viscous damping determine the design displacement
spectrum, reducing the displacement response spectrum by the factor :

10
5
eq

=
+
(3.10)
7. Determine the SDOF substitute structure effective period T
eff
, as the period
corresponding to
d
.
8. Determine the effective stiffness k
eff
associated to the SDOF system maximum
response:

2 2
4 /
eff eff eff
k m T = ; 2
eff
eff
eff
m
T
k
= (3.11), (3.12)
9. Determine the system base shear as
b eff d
V k = (Figure 3.2)


Figure 3.2 Base Shear estimate

10. The base shear is distributed as design forces at the different in proportion to the
inelastic displacement.

1
b i i
i n
i i
i
V m
F
m
=

(3.13)
k
eff
= V
b
/
d

Displacement (m)
Base
Shear
(kN)





FBD AND DDBD PROCEDURES
23
3.3 Considerations on equivalent viscous damping
The equivalent viscous damping approach was first proposed by Jacobsen (1930, 1960),
who considered the steady state response of SDOF non linear systems under an
harmonic load and related the equivalent viscous damping to the ratio between the
hysteretic and elastic energy. If a SDOF non linear system subjected to an harmonic load
is considered, it is possible to follow the Jacobsen approach. By assuming a system
response characterized by an elastic force
el
F k u =
and a damping force F
d
which can
be written as
d d
F c u = (as it is in the viscous damping) and considering a displacement
0
sin u u t = , it follows:
0
sin cos
el d d d
F F F ku c u ku t c u t = + = + = + (3.14)
The energy dissipated in one cycle is:
( )
2
2
0
T
dissipated
T
du
E F t dt c u
dt


+
= =

(3.15)
The elastic energy at maximum response is:
2
0
1
2
elastic
E ku =
(3.16)
The ratio between dissipated and elastic energy is:
2
dissipated
d
elastic
E
c
E k

=
(3.17)
Considering that
2
d eq critic eq
c c m = =
the equivalent viscous damping is:
1
4
dissipated
eq
elastic
E
E

=
(3.18)
Considering this approach to describe the system behavior under an earthquake type
excitation leads to underestimate the system response. In fact this approach
overestimates the system damping during an earthquake because it does not consider
the transient response and it is based on an harmonic excitation.
Different equations exist (as reported in Blandon 2005) which relate the equivalent
viscous damping to the system ductility and the post yield stiffness. The equation adopted
in this research is the one proposed by Grant et al. (2004) which relates the equivalent





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
24
viscous damping to the target system displacement ductility (

=
d
/
y
) and the substitute
structure effective period (T
eff
):
( )
1 1
0.05 a 1 1
eq d b
eff
T c

| |
| |
|
= + +
|
|
+ \
\
(3.19)
The procedure adopted by the authors to calibrate the parameters (a, b, c, d) is based on
the force displacement response of SDOF systems subjected to a ductility range from 2
to 6 and with an effective period ranging from 0.5 s to 4 s.
The FBD and DDBD procedures will be compared by means of non linear analyses of
SDOF systems whose non linear behavior is governed by the Takeda hysteretic rule (as
reported in Carr 2006), which well describes reinforced concrete behavior, and the
equivalent viscous damping parameters adopted (Eqn. 3.19) have been calibrated (Grant
et al. 2004) for two sets of Takeda model parameters (Takeda fat and Takeda narrow
model) as it is in Figure 3.3.

y u
Fy
F
ki
rki
p
p
ku=ki

res

a b c d
Takeda fat (=0.3; =0.6; r=0.05)
0.249 0.527 0.761 3.250
Takeda narrow (=0.5; =0; r=0.05)

0.183 0.588 0.848 3.607
Figure 3.3 Equivalent viscous damping parameters for Takeda model

The elastic component of the equivalent viscous damping is related to the secant stiffness
at maximum displacement, therefore this value has to be adjusted (Grant et al. 2004) to
ensure compatibility between substitute and real structure in the non linear analyses.
Adopting an initial stiffness or a tangent stiffness damping for the time history analyses,
the elastic component of the equivalent viscous damping in the substitute structure has to
be corrected (Grant et al. 2004):





FBD AND DDBD PROCEDURES
25
( )
tang
1 1
a 1 1
init e f
eq el el d b
e
T c

| |
| |
= + + + |
|
|
+
\
\
(3.20)
Where
init
el
and
tang
el
refer to how the damping matrix is computed to solve the
equations of motion; being the former value associated to a damping matrix proportional
to the initial stiffness matrix and the latter to a damping matrix proportional to the tangent
stiffness matrix. The parameters of to the two Takeda models are shown in Table 3.1.
Table 3.1. Takeda parameters for analyses
a b c d e f
Takeda fat (=0.3; =0.6; r=0.05)
0.305 0.492 0.790 4.463 0.312 -0.313
Takeda narrow (=0.5; =0; r=0.05)
0.215 0.642 0.824 6.444 0.340 -0.378

Referring to Figure 3.3 and Figure 3.4, it is possible to determine the equivalent viscous
damping with the Jacobsen approach for the steady state response of the Takeda model,
assuming the same ductility is reached for both directions of excitation.

F
Energy
dissipated
Energy
elastic
A
B
C
y u
p
p
C'
B'
A'
O

Figure 3.4 Takeda steady state response for Jacobsen approach

The comparison between elastic energy and the dissipated energy of Eqn. 3.18 leads to
the expression:






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
26
( ) ( )
2 2
1
2 2 2 2
2
hy



(
= + + + +

(3.21)
Where ( ) 1 1 r

= +

Figure 3.5 shows the comparison between the hysteretic portion of the equivalent viscous
damping equation with the Jacobsen approach according to the expression just
evaluated. It is clear how the Jacobsen approach overestimates the hysteretic damping.

Eq. viscous damping - Takeda hysteretic model
0.00
0.05
0.10
0.15
0.20
0.25
0.30
1.0 2.0 3.0 4.0 5.0 6.0
Ductility ()
E
q
.

v
i
s
c
o
u
s

d
a
m
p
i
n
g
Jacobsen: alfa=0.3; beta=0.6; r=0.05
Grant: alfa=0.3; beta=0.6; r=0.05
Jacobsen: alfa=0.5; beta=0; r=0.05
Grant: alfa=0.5; beta=0; r=0.05

Figure 3.5 Hysteretic damping: Jacobsen vs Grant (Teff=1s)

Considering the methodology used by Grant et al. (2004) to calibrate the equivalent
viscous damping for DDBD, it is possible to note how the calibration followed the
definition of an hysteretic model related to Force-Displacement relationship while in this
research the Takeda model is used to describe the Moment-Curvature relationship. To
compare the results obtained by the FBD and the DDBD procedures, adopting for the
latter the hysteretic damping equations proposed, the relationship between , and r
parameters for the Force-Displacement and Moment-Curvature Takeda model have to be
defined (Figure 3.6). This will be done in the following sub-paragraphs.






FBD AND DDBD PROCEDURES
27
y u
Fy
F
ki
rki
p
p
ku=ki

res

y u
My
M
k'i
r'k'i
p
'p
k'u=k'i
'
res

Figure 3.6 Force-displacement and moment-curvature Takeda model parameters.

3.3.1 Relationship between displacement (

) and curvature (

) ductility
The relationships defined in this sub-paragraph and in the following ones have been
obtained considering a plastic hinge region of length L
p
with constant plastic curvature
p

located at the element ends (Figure 3.7) as it is in the finite element program Ruaumoko
(Carr 2006) adopted in the non linear analyses.

p y
H
Lp

Figure 3.7 Inelastic curvature distribution

With these considerations and adopting an elasto-plastic behavior, the relationship
between the displacement and curvature ductility is
1
1
3
p
L
H

= +
(3.22)





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
28
In the general case, considering the post yield stiffness coefficient r (defined as the post
yield to elastic stiffness ratio), the displacement and curvature ductility relationships are:
( )
1
1
' 3 1 '
p
L
r r
H

= +
+
;
( ) ( ) 1 ' 3 1 ' 1
p
L
r r
H

(
= + +
(

(3.23), (3.24)

3.3.2 Relationship between r and r
Referring to Figure 3.8, at the yielding point:
'
y eff y i
y
EI k
F
H H

= =
;
2
3
y
y
H
=
(3.25), (3.26)

y u
Fy
F
ki
rki

y u
My
M
k'i
r'k'i

Figure 3.8 Evaluation of r-r relationship

At the target displacement:
( )
' '
y i u y
u
u
M r k
M
F
H H
+
= = ;
( )
u y u y p
L H = + (3.27), (3.28)
From the definition of r:
( )
( )
2
' '
' '
i
y u y y
u y
i
i
u y p y u y p y
r k
F F
F F
r k
H
r k
L H L H


+

= = =
+
(3.29)
Therefore:
3
2 2
' ' 1 ' '
'
3 3
i i
p i p eff p
r k r k H H
r r
L H k L H EI L
= = =

,
3
'
p
L
r r
H
= (3.30), (3.31)





FBD AND DDBD PROCEDURES
29
3.3.3 Relationship between and
Referring to Figure 3.9, from F- considerations:
( )
( )
( )
2
2
'
3 3
y u y a
u
res u y u y p
u
r
F H
L H H
k

+
= = + (3.32)

y u
Fy
F
ki
rki
ku=ki

res

y u
My
M
k'i
r'k'i
k'u=k'i
'
res

Figure 3.9 Evaluation of - relationship

From M- considerations:
( )
( ) ( )
'
'
b
res p res p u y u y
L H L H r


(
= = +

(3.33)
Equating
( ) ( ) a b
res res
= leads to:
( ) ( ) ( )
( )
'
'
'
3 3
y u y
u y u y y u y
p p
r
H H
r
L L

+
+ = + (3.34)
Which gives:
( )
( )
'
1 1 ' 1
3 3
1 ' 1
p p
H H
r
L L
r

(
+ + +

=
+
(3.35)





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
30
Therefore:
( ) ( )
( )
ln 1 1 ' 1 ln 1 ' 1
3 3
'
ln
p p
H H
r r
L L

(
( (
+ + + +
(

(

=
(3.36)
( ) ( ) ( )
( )
'
ln 3 ln 1 ' 1 ln 1 ' 1 1
3
ln
p
p
H
L H r r
L

(
( (
+ + + +
(

(

= (3.37)
It is observed how - relationship is governed by the target ductility.

3.3.4 Relationship between and
Referring to Figure 3.10, from F- considerations:
( )
( ) ( )
( ) ( ) ( )
( )( )
2
1
= 1
= 1
3
a
c u u y y u
y y u y p
y
p u y
L H
H
L H


= = + =
+ + =
+
(3.38)
( ) ( )( )
1
c y i c y y i u y
F F rk F rk = + = + (3.39)
y u
Fy
F
ki
rki
p
p
c
c

y u
My
M
k'i
r'k'i
p
'p
c
c

Figure 3.10 evaluation of - relationship

From M- considerations:
( ) ( ) ' ' 1 '
c u u y y u
= = + (3.40)





FBD AND DDBD PROCEDURES
31
( ) ( )( )
' ' ' ' 1 '
c y i c y y i u y
M M r k M r k = + = + (3.41)
( )
( ) ( )( )
2
1 '
3
b y
c y c y p u y p
H
L H L H

= + = + (3.42)
Equating
( ) ( ) a b
c c
= leads to
' = (3.43)

3.4 Displacement response spectrum dependence from damping
According to the current version of Eurocode 8 (2004), the displacement spectrum
amplification for damping values different from 5% can be taken into account with the
factor defined before (Eqn. 3.10). Phase 1 (Chapter 5) of this research deals with the
comparison between the FBD and the DDBD procedures; to reduce uncertainties in the
displacement spectrum reduction due to the damping level (in the DDBD procedure), the
factor has been calibrated for the records adopted. This has been done with a least
square procedure applied in the period range 0 4 s to the mean displacement spectrum.
The new spectrum reduction factor adopted is (Figure 3.11)
( )
7.8/ 2.8
eq
= + (3.44)
=SD(x)/SD(5%)
0.5
0.6
0.7
0.8
0.9
1
1.1
1.2
1.3
1.4
2 3 4 5 6 7 8 9 10 11 12 13 14 15
Equivalent viscous damping (%)
S
D
(
x
)
/
S
D
(
5
%
)
Computed Data
g = 7.8
g = 7
g = 10 (EC8)
0.5
( )
(5%) ( 5)
SD g
SD g

| |
=
|
+
\

Figure 3.11 Spectrum reduction factor as function of the damping value

33
4. PRECAST STRUCTURES CONSIDERATIONS


In this chapter the force reduction factor (q-factor) for precast concrete structures is
defined. Specific considerations on the application of DDBD procedure to the precast
concrete structures considered are made: compared to traditional reinforced concrete
structures the typology under exam presents lower displacement ductility demand due to
the higher interstory height (Chapter 2.2); this suggests to check the implications of the
equivalent viscous damping equations, which have been derived for larger displacement
ductility values.

4.1 Definition of the q-factor for precast concrete buildings
According to Eurocode 8 (2004) 5.1.2 it is possible to define two possible structural
types for the precast concrete structures analyzed in this research, depending on the type
of connections used:
1. Frame system: structural system in which both the vertical and lateral loads are
mainly resisted by spatial frames whose shear resistance at the building base
exceeds 65% of the total shear resistance of the whole structural system
2. Inverted pendulum system: system in which 50% or more of the mass is in the
upper third of the height of the structure, or in which the dissipation of energy
takes place mainly at the base of a single building element. NOTE One-storey
frames with column tops connected along both main directions of the building and
with the value of the column normalized axial load
d
nowhere exceeding 0.3, do
not belong to this category.
Thus, from the previous note, considering that the columns are connected in one direction
directly by means of L-beams or inverted T-beams, and in the other direction indirectly by
means of double-T or Omega beams, it is possible to consider the precast system as a
frame system.

The force reduction factors for horizontal seismic actions are defined in Eurocode 8
5.2.2.2. In the case under exam the q-factor is:
0
1.5
w
q q k = (4.1)





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
34
Where
0
q is taken from Table 4.1:
Table 4.1. Eurocode 8 q0 values
Ductility Class Medium
(DCM)
Ductility Class High
(DCH)
Frame System
1
3.0 /
u

1
4.5 /
u

Inverted Pendulum System 1.5 2

with
1
/ 1.1
u
= for one story frame system buildings and
w
k = 1 for frame systems.
In the particular case of precast concrete structures q
p
=q x k
p
, where k
p
is a reduction
factor depending on the energy dissipation capacity of the precast structure (k
p
=1 if the
connections are designed on the basis of the capacity design rules otherwise k
p
=0.5)
In this research a one story frame system with a ductility class DCM has been
considered, which leads to q = 3.3.
The choice to assign a ductility class DCM instead of DCH can be justified by the
presence of cantilever pinned top columns as primary elements to dissipate energy and
accommodate inelastic displacements, with the development of a plastic hinge at their
base; therefore assigning a level of ductility high will lead to a failure in the inelastic
displacement control due to the higher inelastic rotation of the column base.
Regarding the connection system, Eurocode 8-1:2004 identifies three different situations:
1. Connections located outside the critical regions which do not affect the energy
dissipation capacity of the structure.
2. Connections located in the critical regions designed to remain elastic and to
relocate the inelastic response in other regions inside the elements.
3. Connections located in the critical regions designed to carry the inelastic
response.
In the present study connections located outside the critical regions will be considered.






PRECAST STRUCTURES CONSIDERATIONS
35
4.2 Precast concrete structures compared to other structures
Considering the typical warehouse precast concrete structures layout (Figure 2.2,
Figure 2.3, Figure 2.4), the story height is 2-3 times bigger than the other reinforced
concrete structures. This leads to a lower amount of ductility demand as it is clear from
the comparison of a precast structure to typical concrete structure with the same column
cross section.
Indicating with the maximum allowed drift, equations 3.6 and 3.7 for precast concrete
structures are:
2.1
y precast
y
precast
B

= ;
2
3
precast y precast
y
H
= ; (4.2), (4.3)
While the maximum displacement and displacement ductility are:
precast
u precast
H = ;
3
2.1
u precast precast precast
y precast y precast
B
H


= =

(4.4), (4.5)
In the case of classic reinforced concrete structures the previous equations become:
2.1
y usual
y
usual
B

= ;
2
3
usual y usual
y
H
=
; (4.6), (4.7)
usual
u usual
H = ;
3
2.1
u usual usual usual
y usual y usual
B
H


= =

(4.8), (4.9)
The ductility ratio is:
precast
usual
usual
precast
H
H

= (4.10)
Thus, if 2 3
precast usual
H H = (as it usually happens), the ductility is
2 3
usual
precast

.
The low amount of ductility required leads to a limit state related to the interstory drift
code control rather than a material strain limit requirement.
The target ductility could be sensibly small compared to usual reinforced concrete
structures, therefore the hysteretic damping equations (Grant et al. 2004) need to be
checked for ductility values less than 2 (the parameters in those equations have been
calibrated for ductility values between 2 and 6). This has been done by means of non





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
36
linear time history analyses on single degree of freedom systems with the fat Takeda
hysteretic model (Grant et al. 2004). Systems compatible with case study 1 (Figure 2.2)
have been chosen so that
y
(evaluated with Eqn. 3.7) is the same of concrete columns
with H = 7.9 m and square cross section size between 60 and 110 cm. The DDBD
procedure has been applied to the systems and the results from the design have been
compared with the results obtained from the non linear time history analyses.
Figure 4.1 compares the equivalent viscous damping computed in the design process
with the one effectively obtained: the damping obtained from the design equations
underestimates the system damping, which leads to a conservative estimation of the
ductility. Figure 4.2 shows the maximum displacement as a function of the longitudinal
reinforcement ratio: a cross section size greater than 80 cm is needed to limit the
interstory drift to 2.5%.

1 1.5 2 2.5 3 3.5
0
2
4
6
8
10
12
14
16
18
Hysteretic Damping Ductility

H
y
s
t
e
r
e
t
i
c

D
a
m
p
i
n
g

(
%
)


Numerical Data
Grant eq
n
T
eff
= 0.5 s
T
eff
= 4.0 s

1 1.2 1.4 1.6 1.8 2
0.5
0.6
0.7
0.8
0.9
1
1.1
1.2
1.3
1.4
1.5
(Target ductility Real ductility) Ratio

target
(


t
a
r
g
e
t
)

/

(


o
b
t
a
i
n
e
d
)

Figure 4.1 Hysteretic damping and target ductility comparisons





PRECAST STRUCTURES CONSIDERATIONS
37

0 200 400 600 800 1000 1200 1400
0.15
0.3
0.45
0.6
Maximum Displacement as a function of yield force
F
y

u
1% 4%


60 cm
2.5% drift

0 200 400 600 800 1000 1200 1400
0.15
0.3
0.45
0.6
Maximum Displacement as a function of yield force
F
y

u
1% 4%


70 cm
2.5% drift

0 200 400 600 800 1000 1200 1400
0.15
0.3
0.45
0.6
Maximum Displacement as a function of yield force
F
y

u
1% 4%


80 cm
2.5% drift

0 200 400 600 800 1000 1200 1400
0.15
0.3
0.45
0.6
Maximum Displacement as a function of yield force
F
y

u
1% 4%


90 cm
2.5% drift

0 200 400 600 800 1000 1200 1400
0.15
0.3
0.45
0.6
Maximum Displacement as a function of yield force
F
y

u
1% 4%


100 cm
2.5% drift

0 200 400 600 800 1000 1200 1400
0.15
0.3
0.45
0.6
Maximum Displacement as a function of yield force
F
y

u
1% 4%


110 cm
2.5% drift

Figure 4.2 Maximum displacement as a function of reinforce ratio.


39
5. PHASE 1: FBD AND DDBD COMPARISON


In this chapter the equations developed in Chapter 3 are used to compare the FBD and
DDBD procedures. Two possible ways of comparison are presented and applied to the
design of Case Study 1. The moment curvature relationship adopted to describe the
column flexural behavior is the same used to calibrate the equivalent viscous damping
equations in the literature, i.e. the fat Takeda model, whose choice is justified by the
usually low axial load on columns of warehouse structures. At the end of the chapter the
DDBD is applied to precast structures with a different moment-curvature relationship
showing the need of a more rigorous calibration of the equivalent viscous damping
equations. Table 5.1 contains Case Study 1 data.
Table 5.1. Case Study 1 data
Geometric data
Column height 7.9 m
Number of columns 56
Total mass 4645000 kg
Tributary column mass (weight) 82946 kg (814 kN)
Material data
Concrete C 40/50
f
ck
40 (MPa)
f
cd
26.5 (MPa)
f
ctm
3.5 (MPa)
f
ctk
2.45 (MPa)

Rd
0.4 (MPa)
E
c
34525 (MPa)
Steel FeB 44k
f
yk
430 (MPa)
f
yd
374 (MPa)

y
0.00209 (MPa)
E
s
206000 (MPa)






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
40
Regarding the DDBD procedure, in this first phase of the research the model adopted to
describe the structural behavior is a fixed end cantilever pin connected to the top beams.
In these conditions the effective height and the effective mass are equal to the height and
mass of the structure (H
eff
= 7.90 m, m
eff
= 814 kN). The target displacement is the one
corresponding to 2.5% drift:
d
= 0.1975 m.

5.1 FBD-DDBD comparison: Procedure 1
The procedure adopted is: design the column cross section with the FBD procedure,
calculate the inelastic displacement as
( ) ( )
3
/ 3
u b eff
q V H EI = and use this value as
the target displacement for the DDBD procedure. The following step is to check the
procedure predictions by means of Non Linear Time History (NLTH) analyses.
To obtain the NLTH model input data according to the FBD:
1. Assume, as it is usually done in practice, that at yield 0.50
eff gross
I I = , so

0
/ 0.5
eff gross
k k = = (5.1)
2. Find F
y
and
y
:
1
u eff u
y
F r k
F
r

=

;
y
y
eff
F
k
=
(5.2), (5.3)
3. Get
( ) 1
u u
eff
y u eff u
k r
F r k


= =

(5.4)
4. Get (Eqn. 3.36), (Eqn. 3.43) and r (Eqn. 3.31).

To obtain he NLTH model input data according to the DDBD:
1. Get
y
associated to the cross section size (Eqn. 3.7).
2. Get /
u y

= (5.5)
3. Find
( ) 1 1
u
y
F
F
r

=
+
with r=0.05 (5.6)
4. Get
y
eff
y
F
k =

;
3
0
3
eff eff
gross gross
k k
H
k EI
= =
(5.7), (5.8)
5. Get (Eqn. 3.36), (Eqn. 3.43) and r (Eqn. 3.31).






PHASE 1: FBD AND DDBD COMPARISON
41
In the case of a cross section size equal to 60 cm the FBD procedure leads to the results
shown in Figure 5.1

0 1 2 3 4
0
0.5
1
1.5
2
EC8 2004 type=1 soil=c PGA=0.5 q=3.30
Period (s)
A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
T=1.89; S=0.14


Design spectrum
Elastic spectrum
FBD result

FBD results:
T
eff
1.39 s
V
b
113 kN
M
u
894kNm
Drift 5.16%
0.37
Figure 5.1 FBD - section size 60 cm

The inelastic displacement obtained from the FBD procedure is 0.41 m, which used as a
target in the DDBD procedure leads to a displacement demand higher than the maximum
values of the design displacement spectrum (Figure 5.2).

0 0.5 1 1.5 2 2.5 3
0
0.1
0.2
0.3
0.4
0.5
0.6
Period (s)
D
i
s
p
l
a
c
e
m
e
n
t

(
m
)
EC8 2004 type=1 soil=c PGA=0.5


Design
Elastic
Target displacement

Figure 5.2 DDBD - section size 60 cm Displacement demand higher than capacity.

The DDBD procedure shows that the target displacement corresponds to a ductility level
greater than the one available, therefore iterations are necessary to reduce the target





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
42
displacement until reaching convergence. Figure 5.3 shows the results of these iterations,
which lead to a lower target displacement corresponding to a structure with as effective
period the displacement spectrum corner period, in this case 2 s.

0 0.5 1 1.5 2 2.5 3
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
Period (s)
D
i
s
p
l
a
c
e
m
e
n
t

(
m
)
EC8 2004 type=1 soil=c PGA=0.5
T=2.00; Sd=0.30


Elastic Spectrum
Design Spectrum
DDBD result

Figure 5.3 DDBD - section size 60 cm converged results

The comparison between FBD and DDBD procedures leads to the following results
obtained varying the column cross section size from 70 cm to 110 cm (Table 5.2 to
Table 5.6).
Table 5.2. Cross section size 70 cm
FBD Takeda model parameters
T
eff
(s) 1.39 M
y
(kNm) 1070
V
b
(kN) 154
0
0.5
M
u
(kNm) 1217 0.47
drift (%) 3.79 0.6
0.20 r 0.015
DDBD Takeda model parameters
T
eff
(s) 1.92 M
y
(kNm) 1977
V
b
(kN) 266
0
0.55
M
u
(kNm) 2098 0.45
drift (%) 3.79 0.6
0.12 r 0.015






PHASE 1: FBD AND DDBD COMPARISON
43
Increasing the cross section size leads to a decrease of the inelastic displacement
according to FBD, while the decrease of the target displacement in the DDBD procedure
leads to a greater demand.
Table 5.3. Cross section size 80 cm
FBD Takeda model parameters
T
eff
(s) 1.06 M
y
(kNm) 1630
V
b
(kN) 229
0
0.5
M
u
(kNm) 1813 0.46
drift (%) 2.92 0.6
0.12 r 0.015
DDBD Takeda model parameters
T
eff
(s) 1.44 M
y
(kNm) 2749
V
b
(kN) 365
0
0.51
M
u
(kNm) 2882 0.44
drift (%) 2.92 0.6
0.08 r 0.015

Table 5.4. Cross section size 90 cm
FBD Takeda model parameters
T
eff
(s) 0.84 M
y
(kNm) 1779
V
b
(kN) 256
0
0.5
M
u
(kNm) 2024 0.47
drift (%) 2.31 0.6
0.07 r 0.015
DDBD Takeda model parameters
T
eff
(s) 1.18 M
y
(kNm) 3269
V
b
(kN) 428
0
0.43
M
u
(kNm) 3381 0.43
drift (%) 2.31 0.6
0.04 r 0.015







DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
44
Table 5.5. Cross section size 100 cm
FBD Takeda model parameters
T
eff
(s) 0.68 M
y
(kNm) 2204
V
b
(kN) 317
0
0.5
M
u
(kNm) 2508 0.47
drift (%) 1.87 0.6
0.05 r 0.015
DDBD Takeda model parameters
T
eff
(s) 0.90 M
y
(kNm) 4549
V
b
(kN) 598
0
0.44
M
u
(kNm) 4727 0.42
drift (%) 1.87 0.6
0.03 r 0.015

Table 5.6. Cross section size 110 cm
FBD Takeda model parameters
T
eff
(s) 0.56 M
y
(kNm) 2461
V
b
(kN) 354
0
0.5
M
u
(kNm) 2800 0.47
drift (%) 1.43 0.6
0.03 r 0.015
DDBD Takeda model parameters
T
eff
(s) 0.64 M
y
(kNm) 7023
V
b
(kN) 903
0
0.5
M
u
(kNm) 7137 0.40
drift (%) 1.43 0.6
0.01 r 0.015

Figure 5.4 summarizes the results obtained from the FBD and DDBD comparison, where
the base moment demand according to the two procedures has been plotted; the DDBD
target displacement has been taken as the inelastic displacement predicted by FBD. In
the dashed lines the moment capacity for longitudinal steel ratio of 1 to 4% is plotted.





PHASE 1: FBD AND DDBD COMPARISON
45
0 1 2 3 4 5 6 7
0
1000
2000
3000
4000
5000
6000
7000
8000
9000
Base Moment
M
o
m
e
n
t

(
k
N
m
)
Column size 1=60, 2=70, 3=80, 4=90, 5=100, 6=110 cm


1%
2%
3%
4%
5
.
1
6
%

d
r
i
f
t
3
.
7
9
%

d
r
i
f
t
2
.
9
2
%

d
r
i
f
t
2
.
3
1
%

d
r
i
f
t
1
.
8
7
%

d
r
i
f
t
1
.
4
3
%

d
r
i
f
t
FBD
DDBD

Figure 5.4 FBD and DDBD procedures comparison results.

According to FBD, discarding the results of the 60 cm cross section due to the high value
of theta (second order effects), it seems possible to design any cross section size greater
than 70 cm. The increase of the cross section size leads to a lower value of the
displacement computed by the FBD procedure and a related increase of the moment
demand. Non Linear Time History (NLTH) analyses, with the Takeda hysteretic model
parameters shown in Table 5.2 to Table 5.6, have been carried out. The results are
shown in Table 5.7 and Figure 5.5.
Table 5.7 FBD and DDBD NLTH analyses comparison (procedure 1).
Section size Procedure Target Drift
Time history results
Roof Drift (%) Residual drift (%)
(cm) (%) Mean Max Mean max
70 FBD 3.79 3.06 4.29 0.18 0.38
DDBD 3.79 2.67 4.10 0.09 0.23
80 FBD 2.92 2.28 3.40 0.17 0.48
DDBD 2.92 1.98 2.77 0.12 0.27
90 FBD 2.31 1.90 3.27 0.23 0.40
DDBD 2.31 1.71 2.48 0.11 0.23
100 FBD 1.87 1.47 2.13 0.17 0.55
DDBD 1.87 1.51 2.18 0.03 0.13
110 FBD 1.43 1.33 2.04 0.10 0.24
DDBD 1.43 1.16 1.58 0.05 0.20





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
46

0 1 2 3 4 5 6 7
0
1
2
3
4
5
Roof drift Mean values
D
r
i
f
t

(
%
)
Column size 1=60, 2=70, 3=80, 4=90, 5=100, 6=110 cm


Target displacement
FBD Drift
FBD Residual drift
DDBD Drift
DDBD Residual drift

0 1 2 3 4 5 6 7
0
1
2
3
4
5
Roof drift Max values
D
r
i
f
t

(
%
)
Column size 1=60, 2=70, 3=80, 4=90, 5=100, 6=110 cm


Target displacement
FBD Drift
FBD Residual drift
DDBD Drift
DDBD Residual drift

Figure 5.5 FBD and DDBD THA mean and maximum values comparison (procedure 1).

The two procedures present comparable results with a slightly better control of the target
displacements for the DDBD procedure. The good agreement between the displacement
predicted by the FBD and the displacement obtained in the NLTH analyses is due to the
column hysteretic model adopted in the latter which reflects exactly the approximation
EI
eff
= 50% EI
gross
made in the design procedure. Considering in the design a model which
reflects the actual displacement at yielding (Priestley 2003) leads to the values in
Table 5.8.
Table 5.8 NLTH analyses results: effective stiffness dependence.
Section size Target Drift EI
eff
= 0.5 EI
gross

y
(Priestley 2003)
Roof Drift (%) Roof Drift (%)
(cm) (%) Mean Max Mean Max
70 3.79 3.06 4.29 3.22 4.39
80 2.92 2.28 3.39 2.79 4.15
90 2.31 1.90 3.27 2.21 3.23
100 1.87 1.47 2.13 1.79 2.43
110 1.43 1.33 2.03 1.66 2.29

It is clear how the choice of the effective stiffness affects the results. As noted before the
effective stiffness in the FBD procedure is reduced usually as a percentage of the gross
section stiffness and for columns with low axial load, as it is in this case study, the
recommended value (Paulay and Priestley 1992) is 0.4 EI
gross
; this leads to a period





PHASE 1: FBD AND DDBD COMPARISON
47
which does not take into account the actual longitudinal reinforcement ratio
l
. The FBD
procedure considers the stiffness as a property of the section, while the system property
that does not change is the yield displacement (Priestley 2003). The effective period
associated to the actual yielding moment (depending on
l
) can be detected by:
1.
y
eff
y
M
EI

= where
y
is taken from Eqn. 3.6
2. 2
eff
m
T
k
= where
3
3
eff
EI
k
H
= (valid for fixed end cantilever)
This leads to the period values shown in Table 5.9:
Table 5.9 Steel ratio-period dependence.

Section
size
Period dependence
Steel ratio -
l
with with

1% 2% 3% 4% EI
gross
0.4 EI
gross

(cm) (s) (s) (s) (s) (s) (s)
60 2.85 2.16 1.90 1.73 1.34 2.11
70 2.00 1.71 1.41 1.30 0.98 1.55
80 1.56 1.27 1.08 0.97 0.75 1.19
90 1.25 0.99 0.87 0.78 0.59 0.94
100 1.05 0.83 0.71 0.62 0.48 0.76
110 0.88 0.69 0.58 0.51 0.40 0.63

It is clear how the steel ratio affects the period and therefore the demand predicted by the
FBD procedure. In this specific case the period evaluated with a reduced stiffness
(0.4 EI
gross
) is close to the period corresponding to 2% steel ratio, which is a ratio
reasonably adopted in design.
The period dependence on the longitudinal reinforcement ratio just shown suggests the
idea of including it in the DDBD procedure. In fact knowing if the moment capacity of a
given cross section size is greater than the moment demand without calculating the
former will improve the DDBD procedure. This could be done with the following procedure
which gives an estimation of the longitudinal steel ratio (
l
) based on the material
characteristics and on the ratio between the effective and the gross section stiffness
obtained from the DDBD procedure.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
48
According to DDBD procedure, once the cross section size has been defined, the
y
is
known (Priestley 2003); after the definition of the target displacement the displacement
ductility

is found and as an intermediate result of the procedure the effective period of


the substitute structure T
eff
. From these values the period T
0
(associated to the secant
stiffness at yield) is:
0
1 ( 1)
eff
r
T T

+
=
(5.9)
Where r is the post yield stiffness ratio. The stiffness at first yield is obtained from T
0
as
2
0
4
y
m
k
T

= (5.10)
Which equated to the force-displacement stiffness of a cantilever column gives:
3
3
y
c y
H k
E I = (5.11)
The ratio between the moment-curvature stiffness at yield and in the un-cracked
conditions (considering only the concrete contribution) is:
( ) ( )
0
/
c y c gross
E I E I = (5.12)
This ratio is obtained from the DDBD procedure and, if related to the longitudinal steel
reinforcement ratio, it will allow to check directly if the cross section chosen is suitable for
the design in terms of Code longitudinal steel ratio limits.
The other way to evaluate E
c
I
y
is with the yield curvature formula (Priestley 2003) already
adopted in the DDBD procedure:
/
y y y
y s
y y y
M M M
EI E B
B f
= = =

(5.13)
Where M
y
comes from the moment-curvature bilinear approximation and is the nominal
moment associated to a yield strain in the extreme tension reinforcement or a strain of
0.002 in the extreme compression fiber, whichever occurs first. indicates the factor
used in the yield curvature formula which varies from the cross section size (2.1 for
rectangular columns).
M
y
is now determined as a function of cross section size B, longitudinal steel ratio
l
and
axial load ratio , with the following approximations: square cross sections with equally





PHASE 1: FBD AND DDBD COMPARISON
49
distributed reinforcement (considering 4, 8 and 12 re-bars); contribution of the
reinforcement in the compressed region neglected; concrete cover (c
cover
) dimension
neglected when compared to the cross section size.
In the case of 4 re-bars (one re-bar at each corner), the vertical force equilibrium leads to:
2 2
0.8
2
l
c c y
f B N xf B f B

= =
(5.14)
Where:
x is the neutral axis depth
d is the effective depth (distance between the extreme compressed fiber and the re-
bars in tension)
N is the axial load

From this the neutral axis depth is:
2
1.6
c l y
c
f f
x B
f
+
=
(5.15)
The yield moment at the cross section centroid is:
2
cov
0.8 0.4
2 2 2
l
y y er c
B B
M f B c xBf x

| | | |
= +
| |
\ \
(5.16)
Substituting the neutral axis depth just obtained gives:
( ) ( )
3 2
2
3
2
2 2
4 2 2 4
2 1
4 2
c l y c l y
l
y y
c
y
c l y l
c
f f f f
B
M f B B B
f
f
B
f f
f


+ + | |
= + =
|
\
(
= +
(
(

(5.17)
The
0
ratio sought is independent from the cross section size:
( )
0 4 3
/
12
/12
y s y c y y
s
c gross y c c
M E B f E I M
E
E I f E E B B

= = = =



( )
2
3
2 1
2
y
s c
l l
c y c
f
E f
E f f

( | |
= + ( |
|

(
\
(5.18)





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
50
In the case of three longitudinal re-bars at each side (8 in total):
( )
2
0
3 3 9
2 1
4 32
y
s c
l l
c y c
f
E f
E f f

( | |
= + ( |
|

(
\
(5.19)
In the case of four longitudinal re-bars at each side (12 in total):
( )
2
0
3 2 4
2 1
3 18
y
s c
l l
c y c
f
E f
E f f

( | |
= + ( |
|

(
\
(5.20)
With these last three equations it is possible to know directly from the DDBD procedure if
the cross section chosen will satisfy longitudinal steel ratio requirements. The results are
summarized in the following Table 5.10

Table 5.10 0 ratio values for 4,8 and 12 re-bars.
f
ck
= 40 MPa f
yk
= 430 MPa f
cd
= 25 MPa - f
yd
= 373 MPa
Square cross section - 4 Re-bars

l
=1% 2% 3% 4%
l
=1% 2% 3% 4%
0.1 0.33 0.48 0.63 0.76 0.28 0.43 0.56 0.69
0.2 0.43 0.57 0.69 0.81 0.35 0.48 0.60 0.70
0.3 0.50 0.62 0.73 0.82 0.39 0.51 0.60 0.69
0.4 0.53 0.63 0.72 0.80 0.41 0.51 0.58 0.65
Square cross section - 8 Re-bars

l
=1% 2% 3% 4%
l
=1% 2% 3% 4%
0.1 0.28 0.40 0.51 0.61 0.24 0.35 0.45 0.54
0.2 0.39 0.50 0.59 0.68 0.31 0.41 0.50 0.58
0.3 0.47 0.56 0.64 0.71 0.36 0.45 0.52 0.58
0.4 0.51 0.58 0.65 0.71 0.39 0.46 0.51 0.56
Square cross section - 12 Re-bars

l
=1% 2% 3% 4%
l
=1% 2% 3% 4%
0.1 0.27 0.37 0.47 0.56 0.22 0.32 0.41 0.49
0.2 0.38 0.47 0.55 0.63 0.30 0.39 0.46 0.53
0.3 0.46 0.53 0.61 0.67 0.35 0.42 0.49 0.54
0.4 0.50 0.56 0.62 0.67 0.38 0.44 0.49 0.53





PHASE 1: FBD AND DDBD COMPARISON
51
5.2 FBD-DDBD comparison: Procedure 2
The second way of comparing FBD and DDBD procedures starts from the DDBD
procedure, adopted to match a target displacement set at 2.5% of the roof height
(damage limitation requirement limit used in this research), and uses this displacement
value to get the q-factor for the FBD procedure:
1. Determine the displacement ductility as
d
y

where
2
3
y
y
H
= ,
2. Get the q-factor from the equivalent displacement approximation q

= , as
stated in the actual Codes.

The comparison procedure leads to the following results varying the cross section sizes
from 60 to 110 cm (Table 5.11 to Table 5.16).

Table 5.11. Cross section size 60 cm
DDBD Takeda model parameters
T
eff
(s) 1.10 M
y
(kNm) 4168
V
b
(kN) 534
0
1.84
M
u
(kNm) 4222 0.39

1.26 0.6
0.09 r 0.015
FBD Takeda model parameters
T
eff
(s) 1.89 M
y
(kNm) 2715
V
b
(kN) 345
0
0.5
M
u
(kNm) 2727 0.37
q 1.26 0.6
0.14 r 0.015






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
52
Table 5.12. Cross section size 70 cm
DDBD Takeda model parameters
T
eff
(s) 1.20 M
y
(kNm) 3466
V
b
(kN) 449
0
0.96
M
u
(kNm) 3548 0.41

1.47 0.6
0.07 r 0.015
FBD Takeda model parameters
T
eff
(s) 1.39 M
y
(kNm) 2665
V
b
(kN) 346
0
0.5
M
u
(kNm) 2733 0.42
q 1.47 0.6
0.09 r 0.015

Table 5.13. Cross section size 80 cm
DDBD Takeda model parameters
T
eff
(s) 1.22 M
y
(kNm) 3319
V
b
(kN) 435
0
0.62
M
u
(kNm) 3433 0.43

1.68 0.6
0.05 r 0.015
FBD Takeda model parameters
T
eff
(s) 1.06 M
y
(kNm) 3031
V
b
(kN) 398
0
0.5
M
u
(kNm) 3143 0.43
q 1.68 0.6
0.06 r 0.015






PHASE 1: FBD AND DDBD COMPARISON
53
Table 5.14. Cross section size 90 cm
DDBD Takeda model parameters
T
eff
(s) 1.24 M
y
(kNm) 3181
V
b
(kN) 421
0
0.42
M
u
(kNm) 3323 0.44

1.89 0.6
0.04 r 0.015
FBD Takeda model parameters
T
eff
(s) 0.84 M
y
(kNm) 3369
V
b
(kN) 447
0
0.5
M
u
(kNm) 3534 0.44
q 1.89 0.6
0.04 r 0.015

Table 5.15. Cross section size 100 cm
DDBD Takeda model parameters
T
eff
(s) 1.26 M
y
(kNm) 3050
V
b
(kN) 407
0
0.29
M
u
(kNm) 3218 0.44

2.10 0.6
0.04 r 0.015
FBD Takeda model parameters
T
eff
(s) 0.68 M
y
(kNm) 3712
V
b
(kN) 499
0
0.5
M
u
(kNm) 3941 0.45
q 2.10 0.6
0.03 r 0.015






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
54
Table 5.16. Cross section size 110 cm
DDBD Takeda model parameters
T
eff
(s) 1.28 M
y
(kNm) 2926
V
b
(kN) 395
0
0.21
M
u
(kNm) 3118 0.45

2.31 0.6
0.03 r 0.015
FBD Takeda model parameters
T
eff
(s) 0.56 M
y
(kNm) 3724
V
b
(kN) 506
0
0.5
M
u
(kNm) 4000 0.45
q 2.31 0.6
0.02 r 0.015

It has to be noted that the inelastic displacement computed with the FBD procedure is
different from the initial target displacement (Table 5.17). This is due to the equal
displacements approximation (q=

). The NLTH analyses, which have been carried out


adopting the Takeda hysteretic model parameters shown in Table 5.11 to Table 5.16,
lead to the results shown in Table 5.17 and Figure 5.6.
The results show how the DDBD procedure succeeds in predicting and controlling the
displacements with a stable trend for both maximum and mean values. As stated before,
the main drawback of the FBD procedure is the inability to determine the target
displacement at the beginning of the analysis.





PHASE 1: FBD AND DDBD COMPARISON
55

Table 5.17. FBD and DDBD NLTH analyses comparison (procedure 2).
Section size Procedure Target Drift Time history results
Roof Drift (%) Residual drift (%)
(cm) (%) Mean Max Mean max
60 DDBD 2.5 1.88 2.49 0.05 0.19
FBD 5.2 3.41 4.25 0.09 0.18
70 DDBD 2.5 1.77 2.61 0.08 0.22
FBD 3.8 2.59 4.00 0.03 0.09
80 DDBD 2.5 1.73 2.56 0.10 0.22
FBD 2.9 2.01 2.80 0.11 0.24
90 DDBD 2.5 1.72 2.50 0.10 0.17
FBD 2.3 1.66 2.38 0.10 0.17
100 DDBD 2.5 1.68 2.43 0.08 0.18
FBD 1.8 1.47 2.25 0.12 0.34
110 DDBD 2.5 1.66 2.35 0.08 0.17
FBD 1.4 1.28 1.90 0.07 0.25

0 1 2 3 4 5 6 7
0
1
2
3
4
5
Roof drift Mean values
D
r
i
f
t

(
%
)
Column size 1=60, 2=70, 3=80, 4=90, 5=100, 6=110 cm


Target displacement
FBD Drift
FBD Residual drift
DDBD Drift
DDBD Residual drift
0 1 2 3 4 5 6 7
0
1
2
3
4
5
Roof drift Max values
D
r
i
f
t

(
%
)
Column size 1=60, 2=70, 3=80, 4=90, 5=100, 6=110 cm


Target displacement
FBD Drift
FBD Residual drift
DDBD Drift
DDBD Residual drift

Figure 5.6 FBD and DDBD NLTH analyses mean and max values comparison (procedure 2).






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
56
5.3 Considerations about the inelastic displacements
As stated in the previous paragraphs the main drawback in the comparison between FBD
and DDBD procedure is the failure of predicting inelastic displacements with the FBD
procedure.
According to Eurocode 8 equation
i d e
d q d = stated in the FBD procedure, it is clear
that, once the system period has been defined, the procedure leads to a displacement
equal to the one obtained using an elastic spectrum, for any value of q-factor (Figure 5.7)

e
Fe/q2
Fe
Fe/q1
Fe/q3
e3 e1 e2
e1 q1
e2 q2
e3 q3

Figure 5.7 FBD inelastic displacement

Other estimations of the inelastic displacement can be found in the literature like the
following approximation (Priestley 1997) of q as function of

, which provides a better


estimation of the inelastic displacements:
( )
0
1 1
1.5
T
q
T

= + , for T<T
0
, where T
0
is the peak spectrum period.

The FBD drawback is to rely on the equal displacement approximation while the
relationship between inelastic displacement and q factor is not so straightforward and
depends on different parameters.
To compare more realistically the two procedures, the inelastic system displacement for
the FBD procedure should be better estimated.





PHASE 1: FBD AND DDBD COMPARISON
57
This could be done multiplying the displacement computed with the elastic spectrum by a
factor depending on the q factor and the elastic period T of the structure, i.e. using an
inelastic displacement spectrum. At this regard Tobolski and Restrepo (2007) adopted a
C

factor to provide an estimate of the inelastic displacement demand:


1
1
a
i
c
e
C
b T


= = +

(5.21)
Where:
e
is the displacement of an elastic SDOF system of period T
i
is the inelastic displacement of SDOF system with initial elastic period T

is the displacement ductility


T is the structural period
a, b, c are parameters depending on the seismic hazard.

According to authors investigation this relationship has been calibrated to represent the
90
th
percentile of the results and it is valid for displacement ductility values up to 8,
structural periods greater than 0.3 seconds and structures located on firm soils.

The idea outlined in this research is to use this factor to relate the elastic displacement
computed with FBD to the inelastic displacement, in order to allow DDBD comparison.
With simple calculations it is possible to obtain from Figure 5.8 the relationship
C q

= (5.22)
Fe
Fy
y e i

Figure 5.8 Force-displacement relationship





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
58
Therefore, once the q factor and the elastic period T of the structure are known, it is
possible to find

and C

solving the following system:


1
1
a
c
C
b T
C q


= +

(5.23)
For the purpose of this research, a C

factor has been defined based on the ground


motion under consideration, in order to better estimate the results obtained following the
FBD procedure. The program Inspect (Carr 2006) has been used, with the Takeda
hysteresis rule, for ductility values 1, 2, 4 and 8.
A modified version of Eqn. 5.21 is proposed to better fit the experimental results:
1
a
d
c
C
b T

= +

(5.24)
The parameters have been calibrated by means of a least square procedure in the period
range 0.2 - 4 s and for ductility values ranging from 2 to 8. The C

factor computed is the


mean C

value of the seven ground motions. The procedure followed in the calibration is
the multidimensional unconstrained nonlinear minimization (Nelder & Mead 1965) which
leads to a=0.18, b=6, c=2.18 and d=0.07 (Figure 5.9).

10
1
10
0
1
2
3
4
5
6
7
8
C

factor data fitting


Period (s)
C



Data

=2
Data fitting

=2
Data

=4
Data fitting

=4
Data

=8
Data fitting

=8

Figure 5.9 C

: data fitting





PHASE 1: FBD AND DDBD COMPARISON
59
The last equation can be applied to get a better estimation of the inelastic displacements
in the FBD procedure. In this way it is possible to overpass the problem of the arbitrary
reduction of the modulus of inertia and the adoption of the equal displacement
approximation.

5.4 DDBD for 2.5% drift
The DDBD procedure has been applied to Case Study 1 for a target drift of 2.5%
(0.1975 m), adopting the equivalent viscous damping equation available in the literature.
In this case the hysteretic model parameters adopted in the NLTH analyses have been
obtained from the cross sections moment curvature analysis (Appendix A) and differs
from the one used to calibrate the equivalent viscous damping equation used in the
design. The second order effects are taken into account multiplying the moment
computed with DDBD by
1
1
(where is the second to first order moment ratio). The
results of the design are shown in Table 5.18 to Table 5.19.

Table 5.18. Cross section size 60-70-80 cm
Section
Size
60 cm
T
eff

V
b
M
u
M
u
+ 2
nd
ord.
(s) (kN) (kNm) (kNm)
1.10 1.26 534 4222 4383
Longitudinal reinforcement ratio required > 4%
Section
Size
70 cm
T
eff

V
b
M
u
M
u
+ 2
nd
ord.
(s) (kN) (kNm) (kNm)
1.18 1.47 464 3669 3852
Longitudinal reinforcement ratio required > 4%
Section
Size
80 cm
T
eff

V
b
M
u
M
u
+ 2
nd
ord.
(s) (kN) (kNm) (kNm)
1.24 1.68 421 3323 3489
Takeda model parameters for
l
close to 3.5%

0
M
y
r
0.44 3489 0.35 0.2 0.0006







DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
60
Table 5.19. Cross section size 90-100-110 cm
Section
Size
90 cm
T
eff

V
b
M
u
M
u
+ 2
nd
ord.
(s) (kN) (kNm) (kNm)
1.26 1.89 407 3218 3379
Takeda model parameters for
l
less than 2.5%

0
M
y
r
0.37 3379 0.35 0.2 0.0006
Section
Size
100 cm
T
eff

V
b
M
u
M
u
+ 2
nd
ord.
(s) (kN) (kNm) (kNm)
1.28 2.10 395 3118 3273
Takeda model parameters for
l
close to 1.5%

0
M
y
r
0.28 3273 0.35 0.2 0.0006
Section
Size
110 cm
T
eff

V
b
M
u
M
u
+ 2
nd
ord.
(s) (kN) (kNm) (kNm)
1.30 2.31 383 3023 3174
Takeda model parameters for
l
less than 1.5%

0
M
y
r
0.28 3174 0.35 0.2 0.0006

The results of the analyses are summarized in the following Figure 5.10, where it is clear
how the DDBD procedure leads to a not proper estimation of the target drift in the case of
hysteretic model parameters adopted in the NLTH analyses different from the ones used
to calibrate the hysteretic damping equation used in the design. This drawback and its
solution will be analyzed in Chapter 8 and Chapter 9.






PHASE 1: FBD AND DDBD COMPARISON
61
0 1 2 3 4 5 6 7
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
5
Roof drift
D
r
i
f
t

(
%
)
Column size 1=60, 2=70, 3=80, 4=90, 5=100, 6=110 cm


Target displacement
DDBD Drift (mean)
DDBD Drift (max)
DDBD Residual drift (mean)
DDBD Residual drift (max)

Figure 5.10 Time history results (2.5% drift)

5.5 Concluding remarks
At the state of the art the comparison between FBD and DDBD procedures is not
straightforward due to a series of drawbacks in the FBD procedure, the main one is the
failure to predict the inelastic displacements. This is due to the way the inelastic
displacements are related to the q-factor and to the choice of the effective stiffness.
The force reduction factor (q-factor) takes into account the system ductility by reducing
the seismic force to apply to the system, it depends on the structural system type and
once the system type is chosen the q-factor is determined. In the code examined
(Eurocode 8, 2004) the inelastic displacement is related directly to the elastic one
(obtained from a reduced spectrum) multiplying the latter by the q-factor (equivalent
displacement approximation). The inelastic displacement depends on the displacement at
yielding which for the FBD procedure is a function of the gross stiffness reduction
adopted.
A possible way to overcome this drawback is to introduce the C

factor in the FBD


procedure; in this way it is possible to predict more reasonable values for the inelastic
displacement starting from the system gross section stiffness. Doing that, both the
procedures lead to comparable values.
Therefore, as stated before, the main drawback of the FBD is focusing the attention on
the force reduction factor disregarding the inelastic displacements, in fact it provides the
inelastic displacement at the end of the procedure (the DDBD chooses the target





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
62
displacement at the beginning of the procedure). It starts with the choice of the q-factor,
gets a seismic force associated to that value, evaluates the inelastic displacement (by
means of C

) associated to that force and checks if the drift (strain) limit is satisfied, if this
does not happen the section size needs to be changed and the procedure repeated.
Therefore this procedure does not provide convergence to the target displacement which
could be associated to a q-factor different from the one chosen; to overcome this
drawback the q-factor should be better related to the structural system ductility.
This drawback is completely overcome in the DDBD procedure, which, starting from the
target displacement, controls the non-linearity of the problem by the equivalent viscous
damping formulation, which leads, at this stage, to good results. The DDBD procedure
focuses the attention to the target displacement. Good agreement between the inelastic
displacement prediction and the displacement computed by means of non linear time
history analyses has been found. In the case the non linear flexural behavior is governed
by a hysteretic model with parameters different from the ones adopted in the design, the
equivalent viscous damping equation needs a re-calibration (Chapter 8 and Chapter 9).


63
6. PHASE 2: DDBD AND SOIL STRUCTURE INTERACTION


In the previous chapter the DDBD procedure has been applied to structures with fixed
end conditions (i.e. considering a rigid foundation). In a real case scenario the non linear
structural response could be affected by the soil-structure interaction (SSI) which in the
case of soils different from a rock site can contribute to the structural response
significantly.
From a input motion perspective, the seismic horizontal displacements in a soil layer vary
toward the free surface of the site, while in the case of rock site conditions they are
considered the same. Therefore a structure with an embedded portion in a soil layer is
subjected to different transversal components excitations leading to a significant rocking
component. Another change in the input motion is due by excavating and placing a rigid
base foundation into the site, leading to an averaging of the horizontal excitation
compared to free field conditions. This geometric averaging of the seismic input motion is
called kinematic interaction. Considering the inertial loads applied to the structure, these
will lead overturning moment and transverse shear acting at the foundation base, causing
deformations in the soil and therefore modifying the motion at the base (inertial
interaction). Other than the change of the input motion, the radiation of energy of the
propagating waves away from the structure laying on a soil layer will result in a damping
increase.
For the purposes of this research, SSI will be taken into account in the DDBD procedure
adding the foundation flexibility and the foundation damping limited to the foundation
rocking motion. Two possible ways are taken into account extending results available in
the literature which consider a single degree of freedom substitute structure obtained by
static condensation without considering the foundation inertia. Both analytical and non
linear analyses are carried out to check the procedure suitability when foundation inertia
is considered.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
64
6.1 First approach: elastic foundation
SSI can be taken into account in a simplified way considering an elastic foundation acting
in series with the super-structure. The elastic foundation stiffness has been evaluated
according to Gazetas (1991) formula for square foundations:
( )
3
0 _
/ 2
3.6
1
f
found
soil
G B
K

(6.1)
Where:
G is the soil shear modulus for small strains and it is evaluated as
2
soil s
v being
soil

the soil density and v
s
the shear wave velocity.
B
f
is the foundation dimension

soil
is the soil Poisson modulus

The radiation damping (c
r
) associated to the foundation rocking (Gazetas 1991) is:
0.05
r found soil LA
c I = v (6.2)
Where:
v
LA
is Lysmers analog wave velocity taken as
( )
3.4
1
LA s

=

v v

found
I is the foundation footprint modulus of inertia
The foundation damping coefficient is
f
= c
r
/(2km)
If subjected to an harmonic excitation of angular frequency the foundation stiffness is
reduced to:
( ) ( )
0 _ found
K K k

= (6.3)
Where
( ) 1 0.26
f
s
B
k

=
v
.
According to Eurocode 8 (2004) in the case of seismic events the effective soil shear
modulus is a function of the soil shear modulus (evaluated for small strains) and the peak
ground acceleration; this reduction overcomes the small strain assumption stated before.
A regression formula of the data available in Eurocode 8 has been evaluated and
adopted in this research as:
4( ) /
1.164
PGA g
red
G G e

= (6.4)





PHASE 2: DDBD AND SOIL STRUCTURE INTERACTION
65
Which, for the actual PGA of 0.575 g adopted in this research, leads to 0.29 G.
With the notation shown in Figure 6.1, neglecting foundation mass and rotational inertia,
the system stiffness is:
( )
2
s
eq
s
K K
K
K K H h

=
+ +
(6.5)
m
H
h
K
A
Ks
q1
r
i
g
i
d

Figure 6.1 1 DOF system approximation

The system roof displacement is the sum of two terms:
( )
system s f s f
H h = + = + + (6.6)
With
s
, due to structural deformation, and
f
, due to foundation rigid rotation (
f
).
Therefore if the target displacement is governed by story drift as it is for the precast
structures considered here, taking into account foundation flexibility, and so the
contribution of the displacement due to foundation rigid rotation, implies a reduction of the
structural deformation allowed with a consequent larger structural demand in order to limit
the structural displacement.
According to Priestley (2007), the system equivalent damping can be determined as
s s f f
eq
s f

+
=
+
(6.7)
which applies for the steady state response of SDOF non linear systems under harmonic
loads.
With these considerations it is possible to apply the DDBD procedure following the
general procedure, although iterations are needed to take into account the new





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
66
foundation dimensions obtained from the design. The procedure starts from fixed end
conditions in order to get a first trial foundation dimensions.
The extended DDBD procedure has been applied to Case Study 1 considering a column
cross section dimension of 90 cm and a flexural behavior described by the fat Takeda
hysteretic rule. The target displacement is the one associated to 2.5% of interstory drift.
Two types of foundations has been considered, acting on the foundation capacity safety
factor (1.5 for flexible foundation and 3.0 for rigid foundation), to check the procedure
dependence from the foundation rotation. The foundation resistance is based on the
Brinch-Hansen Method (Eurocode 7, 2004); this choice does not interfere with the SSI
procedure validation.

Based on the previous considerations the results of the DDBD procedure for fixed end
conditions and considering SSI are shown in the following Table 6.1: the results obtained
are essentially the same for the two types of foundations and, as expected, imply a
greater structural moment demand compared to fixed end conditions.
Table 6.1. DDBD results for fixed end conditions
DDBD results fixed end conditions
V
d
= 435 kN M
structure
= 3433 kNm M
foundation
= 4215 kNm
DDBD results SSI (flexible foundation B
f
=5.0 m)
V
d
= 452 kN M
structure
= 3575 kNm M
foundation
= 4379 kNm
DDBD results SSI (rigid foundation B
f
=5.7 m)
V
d
= 452 kN M
structure
= 3572 kNm M
foundation
= 4379 kNm






PHASE 2: DDBD AND SOIL STRUCTURE INTERACTION
67
6.2 Second approach: inelastic foundation
To improve the previous approach is possible to take into account the non linear
foundation stiffness and foundation damping change associated to the foundation
rotation, as suggested by Paolucci et al. (2009):
0 _
1
1 a
found
m
K K

=
+
(6.8)
( ) ( )
_ min _ max _ min
1 exp
f f f f
= + (

(6.9)

These relationships have been calibrated (Paolucci et al., 2009) by means of non linear
analyses of a macro-element (i.e. a non linear spring which describes the moment
rotation hysteretic behavior of the foundation based on the soil characteristics) and the
parameters shown before (a, , m,
f_max
and
f_min
) have been found dependent on the
vertical safety factor (i.e. the ratio between the gravity load and the foundation ultimate
capacity under a concentric vertical load). Those parameters are a = 800, m = 0.82,

min
= 3.5%,
max
= 35%, = 138.8 for the flexible foundation and a = 550, m = 0.95,

min
= 3.5%,
max
= 35%, = 138.8 for the rigid one.

Paolucci et al. (2009) suggested to follow the procedure proposed by Wolf (1985), who
studied the dynamic steady state response of the single degree of freedom (SDOF)
system in Figure 6.1 under an harmonic ground excitation.
The system damping is evaluated as
( )
2
eq eq
eq s f
s
K K H h
K K


+
= + (6.10)
Which, in the case of a generalized SDOF system, is equivalent to Eqn. 6.7.
The seismic response of the coupled system is obtained from the equivalent SDOF
system of stiffness Eqn. 6.5 and damping Eqn. 6.10 resting on rigid ground and subjected
to an effective ground displacement amplitude obtained reducing the input motion
amplitude by
2 2
/
system structure
. With this last input motion reduction, the displacement
amplitude obtained from the equivalent SDOF system gives directly the displacement
associated to structural distortion.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
68
The DDBD procedure applies directly to the system, iterating until convergence on the
foundation rotation, with the results summarized in Table 6.2. Compared with the
previous approach results (Table 6.1), the structural demand predicted by this approach
gives different values in the case of flexible and rigid foundations.
Table 6.2. Second approach DDBD results
DDBD results Flexible foundation Rigid foundation
M
deman_structure
fixed-end 3433 kNm
M
deman_structure
3458 kNm 4022 kNm
Foundation rotation 0.0081 rad 0.0010 rad
Rigid Rotation Displacement 0.0638 m 0.0081 m
Column displacement 0.1337 m 0.1894 m

The procedure has been validated with non linear time history (NLTH) analyses with the
program Ruaumoko (Carr 2006). The column inelastic moment-curvature relationship is
described by the Takeda hysteresis rule, while the moment-rotation component of the
foundation is described by the Bouc-Wen model used to simulate the foundation macro-
element hysteresis.
The hysteretic rule is governed by the following parameters and equations:
Parameters:
A1 A2 A3 A4 A5 N D3 D4 D5
Where:
A1 Loop Fatness parameter ( 0.1 to 0.9)
A2 Loop Pinching parameter (-0.9 to 0.9)
A3 Stiffness parameter (usually 1.0)
A4 Degradation parameter (usually 1.0)
A5 Strength parameter (usually 1.0)
n Power Factor, Controls Abruptness (1 to 3, usually 1)
D3 Strength Degradation parameter (0.0 to 0.1) (0 no degradation)
D4 Loop Size Degradation (0.0 to 0.2) (0 no degradation)
D5 Stiffness Degradation (0.0 to 0.2) (0 no degradation)





PHASE 2: DDBD AND SOIL STRUCTURE INTERACTION
69
Equations:
1 1 2
found n
found
found
z
B z A A
z

| |
|
= +
|
\

with z
(t=0)
=0 (6.11)
( )
_
3 4 1
5
found
found y
A A B
z
A

(6.12)
z z z t = +
(6.13)
z z t =
(6.14)
( )
0_
1
found y
M r K r M z = +
(6.15)
( )
tangent 0_
1
y found
z
K r r K

(
= +
(


(6.16)
The output of the foundation macro-element has been used to calibrate the hysteretic
model (Figure 6.2):

0.2 0.15 0.1 0.05 0 0.05 0.1 0.15 0.2
2000
1500
1000
500
0
500
1000
1500
2000
MomentRotation Comparison
Rotation (rad)
M
o
m
e
n
t

(
k
N
m
)


FE Model
Macroelement Model

Figure 6.2 Macro-element and FE model comparison

Figure 6.3 shows the hysteresis model parameters after calibration and the cyclic
behavior of the two cases, compared to the backbone curve.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
70

Flexible foundation Rigid Foundation
0.2 0.15 0.1 0.05 0 0.05 0.1 0.15 0.2
1.5
1
0.5
0
0.5
1
1.5
x 10
4
MomentRotation Comparison
Rotation (rad)
M
o
m
e
n
t

(
k
N
m
)


BoucWen model
Backbone Curve

0.2 0.15 0.1 0.05 0 0.05 0.1 0.15 0.2
1.5
1
0.5
0
0.5
1
1.5
x 10
4
MomentRotation Comparison
Rotation (rad)
M
o
m
e
n
t

(
k
N
m
)


BoucWen model
Backbone Curve

A1 = 0.7 A2 = 0.4 A3 = 1 A1 = 0.6 A2 = 0 A3 = 1
A4 = 1 A5 = 1 N = 1 A4 = 1 A5 = 1 N = 1
D3 = 0 D4 = 0 D5=1.3E-06 D3 = 0 D4 = 0 D5=1.5E-06
Figure 6.3 Macro-element and FE model after calibration.

To take into account the different foundation hysteretic damping (with a Jacobsen
approach) resulting from a hysteretic model different from the macroelement, a new
hysteretic damping definition has been evaluated (Figure 6.4) with a least square
procedure and adopted in the design procedure.

Flexible foundation Rigid Foundation
0 0.005 0.01 0.015 0.02
0
10
20
30
40
Equivalent Viscous Damping Comparison
Rotation (rad)
D
a
m
p
i
n
g

(
%
)


Bouc-Wen model
Bouc-Wen LS
Macroelement
0 0.005 0.01 0.015 0.02
0
10
20
30
40
Equivalent Viscous Damping Comparison
Rotation (rad)
D
a
m
p
i
n
g

(
%
)


Bouc-Wen model
Bouc-Wen LS
Macroelement
Figure 6.4 - Equivalent viscous damping (Jacobsen approach) of the Bouc-Wen hysteretic model





PHASE 2: DDBD AND SOIL STRUCTURE INTERACTION
71
Table 6.3 shows the results of the NLTH analyses: the term mean value refers to the
mean of the time histories maximum values and the term max value to the maximum
value of the time histories maximum values.
Table 6.3. NLTH analyses results compared to SSI-DDBD.

SSI DDBD
NLTH analyses
Mean values Max values
Flexible foundation
Roof Drift (%) 2.5 2.16 2.70
M
structure
(kNm) 3458 3492 3546
M
found
(kNm) 4246 4223 4334
Foundation rotation (rad) 0.0081 0.0042 0.0046
Rigid Rotation Displacement (m) 0.0638 0.0332 0.0363
Column displacement (m) 0.1337 0.1374 0.1770
Residual displacement (m) 0.0102 0.0430
Rigid foundation
Roof Drift (%) 2.5 2.14 2.63
M
structure
(kNm) 4022 3971 4037
M
found
(kNm) 4938 5002 5150
Foundation rotation (rad) 0.0010 0.0011 0.0014
Rigid Rotation Displacement (m) 0.0081 0.0087 0.0111
Column displacement (m) 0.1894 0.1604 0.1580
Residual displacement (m) 0.0070 0.0400

In both cases, flexible and rigid foundation, the design procedure leads to a fairly good
prediction of the maximum displacements although the composition of the overall
displacement (i.e. structural deformation plus rigid foundation component) is not predicted
correctly especially in the flexible foundation case where the design procedure
underestimated the column structural response and overestimated the foundation
rotation; this can be associated to the Jacobsen approach adopted to evaluate the
foundation damping and to the way of computing the system damping based on the
structural steady state response under harmonic loads which lead to a damping
overprediction.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
72
The previous results have been obtained from NLTH analyses neglecting the foundation
inertia according to Wolf (1985) approach. The following Table 6.4 summarizes the finite
element results considering foundation mass and rotational inertia.

Table 6.4. NLTH analyses results compared to SSI-DDBD with foundation inertia.

SSI DDBD
NLTH analyses
Mean values Max values
Flexible foundation
Roof Drift (%) 2.5 2.20 2.83
M
structure
(kNm) 3458 3497 3555
M
found
(kNm) 4246 4182 4291
Foundation rotation (rad) 0.0081 0.0037 0.0038
Rigid Rotation Displacement (m) 0.0638 0.0292 0.0300
Column displacement (m) 0.1337 0.1446 0.1936
Residual displacement (m) 0.0110 0.0390
Rigid foundation
Roof Drift (%) 2.5 2.33 2.77
M
structure
(kNm) 4022 3962 4003
M
found
(kNm) 4938 5054 5134
Foundation rotation (rad) 0.0010 0.0010 0.0011
Rigid Rotation Displacement (m) 0.0081 0.0079 0.0087
Column displacement (m) 0.1894 0.1761 0.2101
Residual displacement (m) 0.015 0.0420

Considering the foundation inertia leads to a reduction in the foundation rotation with an
associated increase in the super-structure demand. This increase leads to a consequent
underestimation of the maximum system displacement. In the following section an
attempt is made to quantify this over-prediction of the SDOF system approach compared
to the 2 DOF system.





PHASE 2: DDBD AND SOIL STRUCTURE INTERACTION
73
6.3 Considering foundation inertia
The main drawbacks of the procedure shown in the previous section are neglecting the
influence of foundation inertia, which can lead to underestimate the system displacement
demand, and considering the system steady state response under a ground harmonic
excitation to evaluate the system equivalent viscous damping.
In this section the results of the 2 DOF system in Figure 6.5 (q
1
is the foundation rotation
around point C and q
2
the horizontal displacement at A) are compared under harmonic
ground excitation to the results of the SDOF system adopted before.

C
Bf
Ks
q2
q1
K
r
i
g
i
d
m
H
M, IG
h
hf
B
A

Figure 6.5 2 DOF system approximation

The dynamic response of the two DOF system is obtained from the following equations.
Considering:
2 2
12
f f
G
h B
I M
+
= , ( )
2 1 A
q q H h = + + v ,
1
2
f
B
h
q = v (6.17), (6.18), (6.19)
The system kinetic energy is:
( ) ( )
2 2 2
1
2
2
2 2
1 2 1 2
1
2
1
2
2 4
G B A
f
G
T I q M m
h
q I M m H h mq mq q H h
( = + + =

( | |
= + + + + + + ( |
|
(
\
v v

(6.20)





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
74
From which the mass matrix is:
( ) ( )
( )
2
2
4
f
G
h
I M m H h m H h
M
m H h m
(
+ + + +
(
=
(
+ (

(6.21)
The potential energy is
2 2
1 2
1 1
2 2
s
V K q K q

= + (6.22)
From which the stiffness matrix is
0
0
s
K
K
K

(
=
(

(6.23)
The damping matrix
0
0
s
c
C
c

(
=
(

(6.24)
The seismic input vector is
( )
eff g
P m r u t = (6.25)
Where r is the influence vector equal to [ ] 0 1
T
r =
In the case of harmonic ground excitation ( ) sin
g
u t the input vector is:
( )
2
sin
eff g
P m r u t = (6.26)
The equations of motion are
eff
MQ CQ KQ P + + =

(6.27)
Considering the homogenous differential equation the eigenvalues are:
( ) ( ) ( )
( )
( )
( )
2 2
2 2 2
2
1, 2
2
2
4
2
s s s
K m K I K m K I K K m I m H h
m I m H h

+ + +
=
+
(6.28)





PHASE 2: DDBD AND SOIL STRUCTURE INTERACTION
75
And the eigenvectors:
( )
01
(1)
2
1 2
01
2
1
u
K I
u
m H h

(
(
=
(
(
+

;
( )
02
(2)
2
2 2
02
2
2
u
K I
u
m H h

(
(
=
(
(
+

(6.29), (6.30)
Where ( )
2
2
2
4
f
G
h
I I M m H h = + + +

Therefore, considering an harmonic ground excitation ( ) sin
g
u t , the equations of
motion solution for a classically damped system leads to:
1
st
mode of vibration
( ) ( ) ( ) ( )
2 2
1 1 1 1 1 1
2 sin
f g
z t z t z t u t + + = (6.31)
2
nd
mode of vibration
( ) ( ) ( ) ( )
2 2
2 2 2 2 2 2
2 sin
s g
z t z t z t u t + + = (6.32)
Where
n
n
n
L
M
= , with
( ) ( ) ( )
1 12 2 22
n T n n
n
L M r m m = = + and
( ) T ( )

n n
n
M M = .
Solving the two differential equations to get z
1
(t) and z
2
(t), the motion of the n-dof is
( )
( ) n
n
z t . The structural distortion is ( )
( )
1
n
n
z t and the foundation rotation is
( )
( )
2
n
n
z t .

In the case of a SDOF system, as in the previous section, under an harmonic ground
excitation ( ) sin
g
u t , the equation of motion is:
( ) ( ) ( ) ( )
2 2
2 sin
eq eq eq g
u t u t u t u t + + = (6.33)
Where
( )
2
eq eq
eq s f
s
K K H h
K K


+
= + ,
( )
2
s
eq
s
K K
K
K K H h

=
+ +
and
2
2
eq
g g
s
u u

= .
As stated before u(t) gives directly the structural distortion.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
76
Figure 6.6 shows the comparison of the 2 DOF and SDOF systems in terms of maximum
system displacement ratio in the two cases. Starting from the results of the SSI-DDBD
procedure; the upper left corner graph shows the influence of the super-structure to
foundation mass ratio as a function of the foundation to super-structure equivalent
viscous damping ratio; the other graphs are obtained changing the foundation dimensions
and the super-structure mass in order to modify the foundation stiffness while keeping the
foundation to super-structure mass ratio.


0 0.5 1 1.5 2
0.95
1
1.05
1.1
1.15
Mass
f
/Mass
s
(

m
a
x

2
d
o
f
)
/
(

m
a
x

1
d
o
f
)
Mass ratio dependence

f
/
s
= 1
f
/
s
= 2
f
/
s
= 3

0 1 2 3 4 5 6 7
0.95
1
1.05
1.1
1.15
K
f
/(K
s
*(H
s
+h
f
)
2
)
(

m
a
x

2
d
o
f
)
/
(

m
a
x

1
d
o
f
)
Foundation dimension dependence - m
s
/M
f
= 0.5

f
/
s
= 1
f
/
s
= 2
f
/
s
= 3

0 1 2 3 4 5 6 7
0.95
1
1.05
1.1
1.15
K
f
/(K
s
*(H
s
+h
f
)
2
)
(

m
a
x

2
d
o
f
)
/
(

m
a
x

1
d
o
f
)
Foundation dimension dependence - m
s
/M
f
= 1.0

f
/
s
= 1
f
/
s
= 2
f
/
s
= 3

0 1 2 3 4 5 6 7
0.95
1
1.05
1.1
1.15
K
f
/(K
s
*(H
s
+h
f
)
2
)
(

m
a
x

2
d
o
f
)
/
(

m
a
x

1
d
o
f
)
Foundation dimension dependence - m
s
/M
f
= 2.0

f
/
s
= 1
f
/
s
= 2
f
/
s
= 3

Figure 6.6 2 DOF 1 DOF comparison

The results, sensitive to the frequency of the harmonic ground excitation here taken as
0.4 Hz corresponding to the period of the equivalent SDOF system, show that the SDOF
approximation could under-predict, up to 15%, the system response.
The under-prediction is sensitive to the foundation to super-structure equivalent viscous
damping ratio and to ( )
2
/
s
K K H h

(
+

.





PHASE 2: DDBD AND SOIL STRUCTURE INTERACTION
77
Therefore, once the SSI-DDBD procedure has been applied, the designer should verify
the two ratios stated before in order to validate the design procedure and check the need
of further non-linear analyses.

6.4 Concluding remarks
In this chapter two methods have been analyzed in order to take into account soil
structure interaction in the DDBD procedure.
The first method considers an elastic foundation with a stiffness reduction to account for
the increase in soil strains due to the seismic event. This method does not deal properly
with foundation non-linearity and is suitable to take into account foundation stiffness in
the system effective stiffness evaluation in the case the displacement component due to
foundation rotation is small compared to the structural distortion.
An extension of the method is considering the actual foundation stiffness and damping
nonlinear dependence on the soil characteristics. The implementation of the SSI in the
DDBD procedure is straightforward, although iterations are needed until convergence on
the foundation rotation. This methodology does not take into account the foundation
inertia which has been shown to play significant role in particular conditions depending on
the foundation and super-structure damping ratio and stiffness ratio. In those situations
the design should be validated with NLTH analyses.


79
7. PHASE 3: DDBD AND BEAM TO COLUMN CONNECTION


In this chapter the influence of the top connection in the DDBD procedure is evaluated.
Considering the top connection leads to a deflected shape related to the first inelastic
mode of vibration (which determines the substitute structure to use in the DDBD
procedure) associated to yielding at the column base and at the beam to column
connection (Figure 7.1). An analytical study is carried out to determine the effective
stiffness and yield displacement dependence on the top connection; these new equations
are implemented in the DDBD procedure and applied to the case studies. The results are
validated by means of non linear time history (NLTH) analyses.

Case 1 Case 2
M
H=Heff
A
C

A
B
C
M
H
Heff

Figure 7.1 Moment distribution with and without considering the top column connection.

7.1 Analytical study
The structural system under exam can be reduced by static condensation to a single
degree of freedom system, therefore the effective mass associated to the first inelastic
mode of vibration to use in the DDBD procedure is the whole system mass.
To study the general effects of the top connection on the system design, different
conditions are analyzed (Figure 7.2):
Case 1: column to beam pinned connection.
Case 2: column to beam general connection with rigid beam.
Case 3: column to beam general connection.
Case 4: column to beam fixed connection.
Case 5: column to beam fixed connection with rigid beam.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
80

A ( shape) B (T shape)
Case 1
H
L/2

H
L

Case 2
k
H
L/2

L
k k
H

Case 3
H
L/2
k

H
L
k k

Case 4
H
L/2

H
L

Case 5
H
L/2

H
L

Figure 7.2 Different approximations adopted in the top connection study.





PHASE 3: DDBD AND BEAM TO COLUMN CONNECTION
81
Based on the previous figure, the initial elastic stiffness (k*) and lateral displacement
(
column
y
) associated to the column base yield are evaluated.

Case 1A and 1B
The elastic stiffness and the yield displacement are obtained directly as:
*
3
3
c
EI
k
H
= ;
2
3
column
y y
H
= (7.1), (7.2)

Case 2A
Considering a system with two degree of freedom (DOF), column top horizontal
displacement and rotation, leads to the following equations
3 2
2
12 6
6 4 0
c c
c c
EI EI
F
H H
EI EI
k
H H

(
( (
= (
( (
(
+
(

(7.3)
From the second equation:
1 6
4
c
c
EI
H EI kH
=
+
(7.4)
Which substituted into the first equation leads to:
3 3
12 6 6
4
c c c
c
EI EI EI
F
H H EI kH
| |
=
|
+
\
(7.5)
Therefore
*
3
12
4
c c
c
EI EI kH
k
H EI kH
+
=
+
(7.6)
The displacement associated to yielding in the top connection, considering the column
connection elastic, is:
4 4
6 6
conn
y conn c c
y y
c c
M
EI kH EI kH
H H
EI EI k

+ +
= = (7.7)
The displacement associated to yielding in the column connection considering the top
connection elastic is obtained from:





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
82
2 2
2
2 6 6
4
column column c c c c
y y
c
EI EI EI EI kH
M
H H H EI kH

+
= + =
+
(7.8)
2
4
6 2
column column c
y y
c c
H EI kH
M
EI EI kH
+
=
+
(7.9)
If
column conn
y y
> the top connection yields before the column. In this case, considering
for sake of simplicity an elasto-plastic connection (a bilinear factor could be add following
the same procedure without altering the results significance) leads to the following new
relationship between lateral force and top displacement:
3
3 3
2
conn
y
c
M
EI
F
H H
= + (7.10)
The displacement associated to yield of the column base is:
2 2
2
6 3 6
column conn column conn
y y y y column
y
c c
M M M
H H
EI EI
| |
= =
|
|
\
(7.11)
This corresponds, considering the right side moment distribution of Figure 7.1, to:
( )( )
C AB BA BC f BC
H H = + + (7.12)
Where the first term is the elastic displacement at B (point of counter flexure) due to the
linear moment distribution from A to B; the second term is the displacement at C due to
the rigid rotation at the point of counter flexure and the third term is the elastic
displacement at B due to the linear moment distribution from C to B.
Therefore at yielding of the base:
( )
( )
2
2
3 2 2 3
column column conn conn
f y f y f y f y column
y f
c c c c
H H M H M H M H H M
H H
EI EI EI EI
| |
= + +
|
|
\

(7.13)
Substituting
column
y
f
column conn
y y
M
H H
M M
=
+
and
conn
y
f
column conn
y y
M
H H H
M M
=
+
:





PHASE 3: DDBD AND BEAM TO COLUMN CONNECTION
83
( ) ( ) ( )
( )
2 2 3
2
2
2 2
2 3
6
2

6 3 6
column column conn conn
y y y y
column
y
column conn
c y y
column conn column conn
y y y y
c c
M M M M
H
EI M M
M M M
H H
EI EI

+
= =
+
| |
= =
|
|
\
(7.14)
Case 2B
The system stiffness is obtained from Case 2A substituting k with 2k. The yield
displacement is obtained substituting
conn
y
M with 2
conn
y
M .
*
3
6 2
2
c c
c
EI EI kH
k
H EI kH
+
=
+
; (7.15)
2 2
3 3 3
column conn column conn
y y y y column
y
c c
M M M
H H
EI EI
| |
= =
|
|
\
. (7.16)

Case 3A
Considering a system with 3 DOF, horizontal displacement and rotation of the column top
and rotation of the beam left side, leads to the following equations
3 2
1
2
2
12 6
0
6 4
0
0
6
0
c c
c c
b
EI EI
H H
F
EI EI
k k
H H
EI
k k
L

(
( ( (
( ( (
+ =
( ( (
( ( (

(

(

(7.17)
From the third equation:
2 1
6
b
kL
EI kL
=

(7.18)
Substituting into the second equation leads to
( )
( )( )
1 2 2
6 6
6 4
b c
b c
H EI kL EI
H EI kL EI kH HLk

=

(7.19)
Which substituted into the first equation leads to





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
84
2
*
3 2
6 6 6
12 3 2
c b c b c
b c b c
EI E I I EI kH EI kL
k
H E I I EI kH EI kL

=

(7.20)
The displacement at yield of top connection, considering the column elastic, is:
2
2
12 3 2
18 3
conn
y conn b c b c
y
b c c
M
E I I EI kH EI kL
H
k E I I EI kL

=

(7.21)
The displacement associated to yielding in the column connection considering the top
connection elastic is:
2 2
2
12 3 2
6 6 3
column column b c b c
y y
c b c b c
H E I I EI kH EI kL
M
EI E I I EI kH EI kL

=

(7.22)
In the case
column conn
y y
> the top connection has yielded and the same results as in
Case 2A apply
2
3 6
column conn
y y column
y
c
M
H
EI
| |
=
|
|
\
(7.23)

Case 3B
As in case 3A substituting I
b
with 2I
b

2
*
3 2
6 12 12
24 6 2
c b c b c
b c b c
EI E I I EI kH EI kL
k
H E I I EI kH EI kL

=

; (7.24)
2 2
3 3
column conn
y y column
y
c
M
H H
EI

= (7.25)

Case 4A
As in case 2A substituting k with
3 6
/ 2
b b
EI EI
L L
=
*
3
6 6
2 3
c c b
c b
EI EI L EI H
k
H EI L EI H
+
=
+
; (7.26)





PHASE 3: DDBD AND BEAM TO COLUMN CONNECTION
85
2
2 3
6 3
column column c b
y y
c b
H EI L EI H
EI L EI H

+
=
+
(7.27)

Case 4B
As in case 2A substituting k with
3 12
2
/ 2
b b
EI EI
L L
=
*
3
3 12
3
c c b
c b
EI EI L EI H
k
H EI L EI H
+
=
+
; (7.28)
2
3
3 6
column column c b
y y
c b
H EI L EI H
EI L EI H

+
=
+
(7.29)

Case 5A and 5B
*
3
12
c
EI
k
H
= ; (7.30)
( )
2
2
/ 2
2
3 6
column column
y y y
H H
= = (7.31)
The results are summarized in Table 7.1.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
86
Table 7.1. Top connection effects on the elastic stiffness and yield displacement
A ( shape) B (T shape)
1
*
3
3
c
EI
k
H
= ;
2
3
column column
y y
H
=

*
3
3
c
EI
k
H
= ;
2
3
column column
y y
H
=
2
*
3
12
4
c c
c
EI EI kH
k
H EI kH
+
=
+


2
3 6
column conn
y y column
y
c
M
H
EI
| |
=
|
|
\

*
3
6 2
2
c c
c
EI EI kH
k
H EI kH
+
=
+


2
3 3
column conn
y y column
y
c
M
H
EI
| |
=
|
|
\

3
*
3
6 6 6
12 3 2
c b c b c
b c b c
EI EI I I kH I kL
k
H EI I I kH I kL

=


2
3 6
column conn
y y column
y
c
M
H
EI
| |
=
|
|
\

*
3
6 12 12
24 6 2
c b c b c
b c b c
EI EI I I kH I kL
k
H EI I I kH I kL

=


2
3 3
column conn
y y column
y
c
M
H
EI
| |
=
|
|
\

4
*
3
6 6
2 3
c c b
c b
EI I L I H
k
H I L I H
+
=
+


2
2 3
6 3
column column c b
y y
c b
H I L I H
I L I H

+
=
+

*
3
3 12
3
c c b
c b
EI I L I H
k
H I L I H
+
=
+


2
3
3 6
column column c b
y y
c b
H I L I H
I L I H

+
=
+

5
*
3
12
c
EI
k
H
= ;
2
6
column column
y y
H
=


*
3
12
c
EI
k
H
= ;
2
6
column column
y y
H
=







PHASE 3: DDBD AND BEAM TO COLUMN CONNECTION
87
In particular the yield displacement of case 2A and 2B are equal to 3A and 3B
respectively, therefore in the DDBD procedure (once it has been verified, as shown
before, that the top connection yields before the column base connection) the substitute
structure will be the same. The substitute structure height will correspond to the point of
counter flexure while the target displacement will be evaluated as it follows.
The effect of the top connection can be easily shown comparing case 1B with case 2B.
Calling
conn
y
column
y
M
M
= :
Case 1A:
2
3
column
y y
H
= (7.32)
Case 2A:
( )
2 2 2
1
3 3 3
column column column
y y y column
y
H H H

= = (7.33)
The displacement and ductility ratio are shown in the following Figure 7.3

2
1
1
y
y
Case
Case


_ 2
_ 1
1
1
Case
Case


0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 0.1 0.2 0.3 0.4 0.5
D
i
s
p
l
a
c
e
m
e
n
t

r
a
t
i
o


0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
0 0.1 0.2 0.3 0.4 0.5
Alpha
D
u
c
t
i
l
i
t
y

r
a
t
i
o


Figure 7.3 Displacement and ductility ratio.

The target displacement associated to the substitute structure has been determined
according to the interstory drift (), as typical of precast concrete structures:
target
H = .





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
88
The displacement due to the rotation of the base plastic hinge is
( )
2
target
1
3
column column
p y y y
H
H H = = = (7.34)
Therefore the target displacement associated to the single DOF system with effective
height the height of counter flexure (
( ) / 1
CF
H H = + ) is
( )
( )
( )
( )
( )
2
2
2
2
2
/ 1
1
3 1

3 1
1
column
y CF
d y p p
column
y
H H
H
H
H

+
= + = + =
+
= +
+
+
(7.35)
The displacement ductility, which has been evaluated for case 3B (Figure 7.2) but can be
easily extended to the other cases, relates directly the substitute structure displacement
to the system one
( )
( )
( )
( )
2 2
2
2
2 2
3 1
/
3 1
1 3 1
column column
y y
d
column
y y
H H
H
H



( (
+
= = + = + ( (
+
+ +
( (

(7.36)
To check the validity of the equations proposed, a parametric study on case 3B
(Figure 7.2) has been carried out and the DDBD procedure results have been compared
to NLTH analyses. The parameters chosen for the study are
( )
/
conn column
y y
M M = ,
/
b c
I I
L H

| | | |
=
| |
\ \
and
c
EI
kH
= , being I
b
and I
c
the modulus of inertia of the beam and the
column respectively.
According to these parameters, the new expression for
conn
y
, which is the displacement
associated to yielding at the beam to column connection being the column base elastic,
is:
24 /( ) 6 / 2 / 24 6 2
3 12 /( ) / 3 12 1
conn conn
y y conn b c b c
y
b c c
M M
EI I HL I k L I k H
H H
k EI I HL I k H k


= =

(7.37)
While
column
y

, which is the displacement associated to yielding at the column base being


the beam to column connection elastic, is:
24 6 2
6 12 6 1
column column
y y
H
M
k



=

(7.38)





PHASE 3: DDBD AND BEAM TO COLUMN CONNECTION
89
The equations in Table 7.1 have been derived considering yielding in the top connection
before yielding of the column base (
conn column
y y
< ); this leads to the choice of parameters
in Table 7.2, where the results of the DDBD procedure to limit the drift to 2.5% are
compared to the NLTH analyses results.
Table 7.2. DDBD and NLTH analyses results comparison.

DDBD
2.5% drift
NLTH Analyses
M
y
(kNm) Mean drift Max drift
0 0.5 0.75 3363 1.87 2.36
1.00 1.90 2.43
1 0.75 1.85 2.32
1.00 1.88 2.38
2 0.75 1.84 2.30
1.00 1.87 2.36
0.2 0.5 0.75 2574 1.98 2.58
1.00 1.98 2.51
1 0.75 1.99 2.55
1.00 1.98 2.56
2 0.75 1.97 2.54
1.00 1.98 2.58
0.4 0.5 0.75 2192 2.00 2.42
1.00 2.05 2.39
1 0.75 1.98 2.46
1.00 2.02 2.41
2 0.75 1.98 2.49
1.00 2.00 2.42
0.6 0.5 0.75 1861 2.09 2.48
1.00 2.15 2.54
1 0.75 2.09 2.41
1.00 2.11 2.51
2 0.75 2.07 2.38
1.00 2.09 2.48

The comparison shows good and stable results between the NLTH analysis and the
DDBD prediction, therefore the equations proposed are suitable for taking into account
the contribution of the beam to column connections in the DDBD procedure.
In this parametric study the DDBD procedure has been applied considering, for sake of
simplicity, one interior column; the joint contribution of the interior and exterior columns
connections can be taken into account considering their mean yielding moment:
0.5
conn conn
y lateral columns y interior columns
conn
mean
lateral columns interior columns
M n M n
M
n n
+
=
+
(7.39)





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
90
7.2 Procedure application to the case studies
To evaluate the procedure potential ability to assess existing structures, the analytical
results found before have been applied to the three case studies shown in Chapter 2.
The DDBD predictions are directly compared with the NLTH analyses.

Case Study 1
The beam is an inverted prestressed T beam. The top connection adopted (Figure 7.4) is
made by two longitudinal bolts with a bent end anchored in the column top. The beam
has two cylindrical hollow holes (made with PVC pipes) where the two bolts are placed.
After placing the bolts a high strength and low shrinkage grout is placed in the holes and
the bolts are post-tensioned to 70-80% of their ultimate resistance. The connection
moment-rotation relationship has been evaluated by means of a moment curvature
analysis of the beam column interface including strain penetration (Figure 7.4).


Top Connection Moment-Rotation
0
20
40
60
80
100
120
140
160
180
200
0 0.05 0.1 0.15 0.2
Rotation (rad)
M
o
m
e
n
t

(
k
N
m
)

Figure 7.4 Case Study 1 beam to column connection.

Table 7.3 shows the DDBD prediction and the NLTH analyses results.





PHASE 3: DDBD AND BEAM TO COLUMN CONNECTION
91

Table 7.3. Case Study 1: DDBD and NLTH analyses results.
Case study data
Base connection M
y
(kNm) 1492
M
u
(kNm) 1615
Top connection M
y
(kNm)
mean value
144
M
u
(kNm)
mean value
158
Height (m) Tributary mass (kg) Cross section size (m)
7.65 m 82950 kg 0.8x0.8
Seismic Zone 1
DDBD prediction NLTH Analysis (mean value)
0.1706 m 0.1794 m

Case Study 2
The beam is an I shape prestressed beam. The top connection adopted (Figure 7.5) is
made by two longitudinal bolts with a bent end anchored in the column top. The bolts are
post-tensioned to 70-80% of their ultimate resistance. The connection moment-rotation
relationship has been evaluated by means of a moment curvature analysis of the beam
column interface including strain penetration (Figure 7.5).


Top Connection Moment-Rotation
0
20
40
60
80
100
120
140
160
180
200
0 0.05 0.1 0.15 0.2
Rotation (rad)
M
o
m
e
n
t

(
k
N
m
)

Figure 7.5 Case Study 2 beam to column connection.






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
92
Table 7.4 shows the DDBD prediction and the NLTH analyses comparison: the top
connection does not affect the results.
Table 7.4. Case Study 2: DDBD and NLTH analyses results.
Case study data
Base connection M
y
(kNm) 765
M
u
(kNm) 803
Top connection M
y
(kNm)
mean value
12
M
u
(kNm)
mean value
13
Height (m) Tributary mass (kg) Cross section size (m)
5.20 m 47200 kg 0.6x0.6
Seismic Zone 2
DDBD prediction NLTH Analysis (mean value)
0.0938 m 0.1010 m

Case Study 3
The beam is a glued laminated timber beam connected to the column top by means of
steel bolts. Two bolts in case study 3a and four bolts in case study 3b (Figure 7.6)

3a 3b
Figure 7.6 Case Study 3a and 3b beam to column connection.






PHASE 3: DDBD AND BEAM TO COLUMN CONNECTION
93
To define the moment rotation relationship, the top connection (Figure 7.7) has been
treated as the dowel connection used in floors with timber beams and composite concrete
topping (Gelfi et al. 2002), the analogy is clear considering half of the connection.

Concrete
column
Dowel
Timber
beam
Wood plate

w
A
B
L
e
L
w
t t
Concrete

Figure 7.7 Connection detail and scheme.

The first approximation is assuming a longitudinal and radial uniform pressure on the
dowel, which corresponds to a distributed load of
//
w
. The connection strength is a
function of dowel length: if the dowel is too short the connection strength is associated to
the wood bearing strength while if the dowel is long enough it is associated to the
development of two plastic hinges in the dowel.
Considering half of the connection for symmetry reasons, the details are (Table 7.5):
Table 7.5. Case Study 3: half connection details.
Dowel diameter 24 mm
Dowel length embedded in the wood L = 118.5 mm
Dowel yield stress
640
y
MPa =
Wood plate thickness 21.5 mm
Wood bearing strength (perpendicular to wood fibers)
3.5
w
MPa

=
Wood bearing strength (parallel to wood fibers)
//
35
w
MPa =

The dowel length associated to a short dowel behavior is:
//
*
1
1
81
3
y
w
L a mm

= + (7.40)





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
94
Where a = 1.5 is the position of the first plastic hinge.
The dowel length associated to a long dowel behavior is:
//
*
2
2
2 189
3
y
w
L a a mm

| |
| = + =
|
\
(7.41)
The actual dowel length (embedded in the wood) is 118.5 mm therefore
*
2
*
1
L L L < < . To
determine the dowel strength (refer to the scheme of Figure 7.7), the portion of the dowel
embedded in the concrete is assumed fixed and the wood plate is not taken into account,
because there is no composite action.
The moment at B, where the shear is zero, is
2
2
//
|

\
|
=
e w
w B
L L
M (7.42)
The moment at A is
( )
B
e
w A
M
L
M =
2
2
//
(7.43)
When the plastic hinge develops the previous formula becomes:
( )
// //
2 2
3
2 2 6
e w e
y w W y
L L L
M


| |
= =
|
\
(7.44)
Which leads to a dowel effective length L
e
=85 mm and to a yield shear of
71.6
half
y y e
V L kN = = (7.45)
The connection stiffness associated (Gelfi et al. 2002) is
3 3
124000 124000 24
21
21.5
4.3
4.3
24
half
kN
K
mm
t


= = =
| | | |
+
+
|
|
\
\
(7.46)
Each dowel goes through two timber-concrete interfaces, therefore the yield strength and
stiffness of a single dowel are 2 143.2
half
y y
V V kN = = , 2 42 /
half
K K kN mm = = .
In the case of two bolts connection, with a dowel distance of 500 mm, the moment and
rotation at yield are respectively 71.6 kNm and 0.00675 rad. The ultimate conditions are
evaluated considering an overstrength factor 1.3 and an ultimate steel strain 0.07.





PHASE 3: DDBD AND BEAM TO COLUMN CONNECTION
95
Analogous considerations apply in the case of four bolts connections. The moment-
rotation relationships are shown for both cases in Figure 7.8.

Case Study 3a Case Study 3b
Top Connection Moment-Rotation
0
20
40
60
80
100
120
140
160
180
200
0 0.05 0.1 0.15 0.2
Rotation (rad)
M
o
m
e
n
t

(
k
N
m
)

Top Connection Moment-Rotation
0
20
40
60
80
100
120
140
160
180
200
0 0.05 0.1 0.15 0.2
Rotation (rad)
M
o
m
e
n
t

(
k
N
m
)

Figure 7.8 Case study 3a and 3b Moment-rotation relationships.

The results of the DDBD procedure and NLTH analyses are shown in Table 7.6 and
Table 7.7.
Table 7.6. Case Study 3a: DDBD and NLTH analyses results.
Case study data
Base connection M
y
(kNm) 985
M
u
(kNm) 1195
Top connection M
y
(kNm)
mean value
36
M
u
(kNm)
mean value
40
Height (m) Tributary mass (kg) Cross section size (m)
5.82 m 50760 kg 0.6x0.6
Seismic Zone 2
DDBD prediction NLTH Analysis (mean value)
0.0902 m 0.0860 m








DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
96
Table 7.7. Case Study 3b: DDBD and NLTH analyses results.
Case study data
Base connection M
y
(kNm) 985
M
u
(kNm) 1195
Top connection M
y
(kNm)
mean value
72
M
u
(kNm)
mean value
80
Height (m) Tributary mass (kg) Cross section size (m)
5.82 m 50760 kg 0.6x0.6
Seismic Zone 2
DDBD prediction NLTH Analysis (mean value)
0.0848 m 0.0832 m

To check if the approximations made in the moment-rotation relationship lead to results in
good agreement with the real connection behavior, two experimental tests were carried
out with the same test setup used for the column to foundation connection evaluation
which will be described in the next chapter.
The specimens simulated a precast concrete column to timber beam connection with two
and four bolts (Figure 7.9) and underwent to cyclic increasing displacements of the top.



Figure 7.9 Four bolts and two bolts timber beam - column connection specimens






PHASE 3: DDBD AND BEAM TO COLUMN CONNECTION
97
The experimental results and the relative comparison with the dowel action approximation
are shown in the following Figure 7.10 and Figure 7.11.

Moment - Rotation relationship (two dowels)
-200
-150
-100
-50
0
50
100
150
200
-0.1 -0.08 -0.05 -0.03 0 0.025 0.05 0.075 0.1
Rotation (rad)
M
o
m
e
n
t

(
k
N
m
)
Experimental data
Approximation

Figure 7.10 Two bolts connection experimental tests

Moment - Rotation relationship (four dowels)
-200
-150
-100
-50
0
50
100
150
200
-0.1 -0.08 -0.05 -0.03 0 0.025 0.05 0.075 0.1
Rotation (rad)
M
o
m
e
n
t

(
k
N
m
)
Experimental data
Approximation

Figure 7.11 Four bolts connection experimental tests






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
98
The dowel approximation gives relatively good results only in terms of ultimate strength.
The initial elastic stiffness is well approximated only for low moment values, after which
the stiffness decreases. Due to permanent deformations in the wood the reloading
stiffness is reduced sensibly. Looking at the test results the dowel approximation works
better in the case of four bolts and can be applied in the DDBD procedure to take into
account the effects of the top connection. In the case of a two bolts connection, the dowel
approximation over-predicts the moment rotation behavior. In this last case the
connection behaves approximately like a pinned connection until reaching a drift of 3-4%.

7.3 Concluding remarks
In this chapter a procedure to take into account the beam-column precast connection has
been developed. The procedure adopts a single DOF substitute structure with height
corresponding to the point of counter flexure of the system and with mass the whole
system mass. The displacement ductility formula proposed relates directly the substitute
structure displacement to the system one. The application of the analytical procedure
gives good results. Another promising way of the procedure application is the assessment
of existing structures. This possibility has been evaluated applying the procedure
successfully to three different case studies.

99
8. PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION


In this chapter the DDBD procedure is applied to systems whose hysteretic behavior is
different from the ones used in the equivalent viscous damping equation calibration.
Experimental tests are carried out to compare different types of precast column to
foundation connections from a seismic and retrofitting point of view and to determine their
hysteretic parameters to use in the equivalent viscous damping equation calibration. A
new and faster calibration algorithm is proposed. The calibration procedure is carried out
for the hysteretic relationships associated to the experimental tests and the results
applied to the DDBD procedure. A new equation to relate the yield curvature to the
column cross section effective depth and the axial load ratio is proposed; this formulation
overcomes the drawback of the equation available in the literature especially when
applied to some foundation to column connections typical of the precast industry.

8.1 Experimental tests
The experimental tests concerned six specimens with different column to foundation
connections and approximately the same maximum bending moment capacity. The entire
specimen tested had a 400x400 mm column cross section and a clear height from the
foundation to the top of 3200 mm.
Specimens CS and PF are representative of a typical cast-in-situ column to foundation
connection and grouted pocket foundation, respectively. Specimens GS4 and GS4B are
both characterized by having four grouted sleeves with different anchorage length of the
26 re-bars in the foundation: the former has straight anchored re-bars, whereas the
latter has 90hooks at the bar ends. Specimen GS8 has 8 grouted sleeves with 22 re-
bars. Specimen GS4U has 300 mm of unbonded length for the 26 re-bars in the column;
this solution allows a simple post seismic event repair as it will be shown later.
The experimental setup adopted is shown in Figure 8.2. For all the tests a 600kN axial
force was first applied by means of two hydraulic jacks. A cyclic horizontal displacement
(Figure 8.1) was then applied at the top of the column by means of a 1000 kN
electromechanical screw jack having a 500 mm maximum stroke.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
100
It is observed that the maximum imposed displacement is equal to 200 mm,
corresponding to a 6.0% drift, much larger than the 2.5% drift commonly accepted for
precast columns under a design seismic event.
-7
-6
-5
-4
-3
-2
-1
0
1
2
3
4
5
6
7
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
Cycles
D
r
i
f
t

(
%
)

Figure 8.1 Loading history

Figure 8.3 to Figure 8.14 show for each test the pictures at drift values of 1%, 2.5% and
5%, the re-bar details and the relationship between:
1. Lateral force and drift.
2. Dimensionless energy: defined as the ratio between the area inside a cycle in the
force - displacement graph and one half of the maximum displacement times the
lateral force at maximum displacement (i.e. the elastic energy).
3. Column base moment and curvature; the latter taken as the average value
between 0-215 mm, 215-430 mm and 430-645 mm (being 0 the foundation to
column interface).
4. Base moment and drift of the experimental test and of the finite element model
which fits the test results.





PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION
101


3
1
B B
1
3
4
2
St. 16/10 L=250
St. 16/10 L=560
8 L=130
St. 8/10 L=166
5
4
3
Section B-B
A A
Section A-A
200
3
2
0
6
0
5
5
5
5
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1
0
2
0
2
0
2
0
2
0
2
0
2
0
2
0
2
0
1
8
6
6
6
4
40
4
0
40
40
4
0
8 24 8
Concrete C35/45 - fck = 35MPa
Steel B450C - f yk = 450MPa - fy,ave = 530 MPa
Sleeves made of corrugated steel prestressing ducts
5 22 L=195
2


Figure 8.2 Experimental test setup





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
102
Cast in Place - CS

2
B B
2
8
7
5
6
4
26 L=171
22 L=195
8 L=130
St. 8/10 L=166
9 18 L=316
2
Section B-B
8
9
4
0
40
9
9 9
CS
8 24 8
8
St. 16/10 L=250
St. 16/10 L=560
6
5
7

CS Force-Drift
-100
-75
-50
-25
0
25
50
75
100
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
Drift (%)
L
a
t
e
r
a
l

f
o
r
c
e

(
k
N
)
Bar Failure
P-
T
h
e
o
r
e
t
i
c
a
l

y
i
e
l
d

Figure 8.3 CS geometry and Force-Drift graph.





PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION
103
0 5 10 15 20
0
0.5
1
1.5
2
2.5
3
3.5
CS Dimensionless Energy
Cycle
E
h
y
s
t
e
r
e
t
i
c

/

E
e
l
a
s
t
i
c

2 1 0 1 2
x 10
4
300
200
100
0
100
200
300
CS MomentCurvature (average values)
Curvature (rad/mm)
M
o
m
e
n
t

(
k
N
m
)


0215 mm
215430 mm
430645 mm

Drift 1.0 % Drift 2.5% Drift 5.0%

6 4 2 0 2 4 6
300
200
100
0
100
200
300
CS MomentDrift Comparison
Drift (%)
B
a
s
e

m
o
m
e
n
t

(
k
N
m
)
FE Model
Experimental test

Takeda parameters r
0.35 0.10 0.005
Figure 8.4 - CS test data and FE model comparison





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
104
Pocket Foundation PF

40
4
0
9
9
2 9
8
9
St. 16/10 L=330 6
10 22 L=72
18 L=371 9
22 L=195
26 L=171
7
8
St. 16/10 L=250
St. 8/10 L=166 2
5
8 L=130 4
8
8
2
4
6
Grout
EMACO S55
B
5
B
7
2
8

PF Force-Drift
-100
-75
-50
-25
0
25
50
75
100
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
Drift (%)
L
a
t
e
r
a
l

f
o
r
c
e

(
k
N
)
T
h
e
o
r
e
t
i
c
a
l

y
i
e
l
dP-


Figure 8.5 PF geometry and Force-Drift graph.





PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION
105
0 5 10 15 20
0
0.5
1
1.5
2
2.5
3
3.5
PF Dimensionless Energy
Cycle
E
h
y
s
t
e
r
e
t
i
c

/

E
e
l
a
s
t
i
c

2 1 0 1 2
x 10
4
300
200
100
0
100
200
300
PF MomentCurvature (average values)
Curvature (rad/mm)
M
o
m
e
n
t

(
k
N
m
)


0215 mm
215430 mm
430645 mm

Drift 1.0 % Drift 2.5% Drift 5.0%


6 4 2 0 2 4 6
300
200
100
0
100
200
300
PF MomentDrift Comparison
Drift (%)
B
a
s
e

m
o
m
e
n
t

(
k
N
m
)
FE Model
Experimental test

Takeda parameters r
0.35 0.10 0.003
Figure 8.6 PF test data and FE model comparison





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
106
Grouted Sleeve Straight re-bars GS4

GS4
St. 16/10 L=250
St. 16/10 L=560 6
22 L=195
26 L=171
18 L=316
8
7
9
40
9
8 L=130
St. 8/10 L=166 2
5
4
2 9
Section B-B
9 9
8
2
4
4
0
8
8
Grout
7
EMACO S55
2
6
5
B B
8

GS4 Force-Drift
-100
-75
-50
-25
0
25
50
75
100
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
Drift (%)
L
a
t
e
r
a
l

f
o
r
c
e

(
k
N
)
T
h
e
o
r
e
t
i
c
a
l

y
i
e
l
d
P-


Figure 8.7 GS4 geometry and Force-Drift graph.






PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION
107
0 5 10 15 20 25
0
0.5
1
1.5
2
2.5
3
3.5
GS4 Dimensionless Energy
Cycle
E
h
y
s
t
e
r
e
t
i
c

/

E
e
l
a
s
t
i
c

2 1 0 1 2
x 10
4
300
200
100
0
100
200
300
GS4 MomentCurvature (average values)
Curvature (rad/mm)
M
o
m
e
n
t

(
k
N
m
)


0215 mm
215430 mm
430645 mm

Drift 1.0 % Drift 2.5% Drift 5.0%

6 4 2 0 2 4 6
300
200
100
0
100
200
300
GS4 MomentDrift Comparison
Drift (%)
B
a
s
e

m
o
m
e
n
t

(
k
N
m
)
FE Model
Experimental test

Takeda parameters r
0.40 0.10 0.0015
Figure 8.8 GS4 test data and FE model comparison





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
108
Grouted Sleeve Bent re-bars GS4B

8
40
4
0
2
4
18 L=316
26 L=171
22 L=195
St. 16/10 L=560
9
7
8
6
St. 8/10 L=166
8 L=130
St. 16/10 L=250
4
5
2
9 9 2
8
Section B-B
GS4B
9 9
8
6
EMACO S55
B
5
B
2
8
7
Grout

GS4B Force-Drift
-100
-75
-50
-25
0
25
50
75
100
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
Drift (%)
L
a
t
e
r
a
l

f
o
r
c
e

(
k
N
)
T
h
e
o
r
e
t
i
c
a
l

y
i
e
l
d
P-


Figure 8.9 GS4B geometry and Force-Drift graph.






PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION
109
0 5 10 15 20 25
0
0.5
1
1.5
2
2.5
3
3.5
GS4B Dimensionless Energy
Cycle
E
h
y
s
t
e
r
e
t
i
c

/

E
e
l
a
s
t
i
c

2 1 0 1 2
x 10
4
300
200
100
0
100
200
300
GS4B MomentCurvature (average values)
Curvature (rad/mm)
M
o
m
e
n
t

(
k
N
m
)


0215 mm
215430 mm
430645 mm

Drift 1.0 % Drift 2.5% Drift 5.0%

6 4 2 0 2 4 6
300
200
100
0
100
200
300
GS4B MomentDrift Comparison
Drift (%)
B
a
s
e

m
o
m
e
n
t

(
k
N
m
)
FE Model
Experimental test

Takeda parameters r
0.45 0.15 0.001
Figure 8.10 GS4B test data and FE model comparison





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
110
Grouted Sleeve 8 re-bars GS8

GS8
EMACO S55
Grout
9 2 9
40
Section B-B
9
8
4
0
8
2
4
9
8
22 L=195 7
22 L=171
18 L=316
8
9
St. 8/10 L=166
8 L=130
St. 16/10 L=560
St. 16/10 L=250
2
5
4
6
B
6
B
7
5
8
2

GS8 Force-Drift
-100
-75
-50
-25
0
25
50
75
100
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
Drift (%)
L
a
t
e
r
a
l

f
o
r
c
e

(
k
N
)
T
h
e
o
r
e
t
i
c
a
l

y
i
e
l
d
P-

Figure 8.11 GS8 geometry and Force-Drift graph.





PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION
111

0 5 10 15 20 25
0
0.5
1
1.5
2
2.5
3
3.5
GS8 Dimensionless Energy
Cycle
E
h
y
s
t
e
r
e
t
i
c

/

E
e
l
a
s
t
i
c

2 1 0 1 2
x 10
4
300
200
100
0
100
200
300
GS8 MomentCurvature (average values)
Curvature (rad/mm)
M
o
m
e
n
t

(
k
N
m
)


0215 mm
215430 mm
430645 mm

Drift 1.0 % Drift 2.5% Drift 5.0%

Figure 8.12 GS8 test data and specimen pictures.

This experimental test exhibited a rigid rotation at the foundation base level due to an
imperfect clamping of specimen GS8 to the reaction frame. This reflects the low level of
energy dissipated and affects the shape of the Force-Drift graph, therefore this test
configuration has been neglected in the finite element modeling.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
112
Grouted Sleeve Unbonded re-bars GS4U

B
5
B
2
8
7
Grout
30 cm unbonded
with duct tape
8
40
4
0
2
4
18 L=316
26 L=171
22 L=195
St. 16/10 L=560
9
7
8
6
St. 8/10 L=166
8 L=130
St. 16/10 L=250
4
5
2
9 9 2
8
Section B-B
GS4U
9 9
8
6
EMACO S55

GS4U Force-Drift
-100
-75
-50
-25
0
25
50
75
100
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
Drift (%)
L
a
t
e
r
a
l

f
o
r
c
e

(
k
N
)
T
h
e
o
r
e
t
i
c
a
l

y
i
e
l
d
P-


Figure 8.13 GS4U geometry and Force-Drift graph.






PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION
113
0 5 10 15 20
0
0.5
1
1.5
2
2.5
3
3.5
GS4U Dimensionless Energy
Cycle
E
h
y
s
t
e
r
e
t
i
c

/

E
e
l
a
s
t
i
c

2 1 0 1 2
x 10
4
300
200
100
0
100
200
300
GS4U MomentCurvature (average values)
Curvature (rad/mm)
M
o
m
e
n
t

(
k
N
m
)


0215 mm
215430 mm
430645 mm

Drift 1.0 % Drift 2.5% Drift 5.0%

6 4 2 0 2 4 6
300
200
100
0
100
200
300
GS4U MomentDrift Comparison
Drift (%)
B
a
s
e

m
o
m
e
n
t

(
k
N
m
)
FE Model
Experimental test

Takeda parameters r
0.35 0 0.01
Figure 8.14 GS4U test data and FE model comparison





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
114
Comparing the energy dissipated (Figure 8.15) by the different types of connections, no
significant differences exist between specimens CS, PF, GS4, and GS4B up to a 2% drift
(cycle 12). Starting from the 2.5% cycles, specimen PF shows a larger energy dissipation,
due to a smaller strength degradation and to a smaller strain localization next to the
column base section. The energy dissipated by specimen GS4U is smaller compared to
the others; this is due to the longitudinal re-bars strain reduction due to the unbonded
length.

0 2 4 6 8 10 12 14 16 18 20
0
0.5
1
1.5
2
2.5
3
Dimensionless Energy
Cycle
E
h
y
s
t
e
r
e
t
i
c

/

E
e
l
a
s
t
i
c


CS PF GS4 GS4B GS4U GS8

Figure 8.15 Dimensionless energy comparison

The high ductility of the grouted sleeve solutions is related to the high confining effect of
the corrugated steel sleeves on the grout contained within. Furthermore, the presence of
a highly confined grout prevents longitudinal re-bars buckling. In such connections the
damage is localized in the 20 mm grout layer existing between the precast column and
the foundation. As a result, very little damage may be observed in the column outside the
base section. Due to the damage localization observed, an easier post-seismic column
repair has to be expected for the grouted sleeve column to foundation connections, with
respect to cast-in-situ or pocket foundation solutions.
To check the retrofit suitability of grouted sleeve solutions another test has been carried
out after repairing the specimen GS4U with the replacement of the damaged grout layer,
as it is shown in the following Figure 8.16.





PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION
115


Figure 8.16 GS4U specimen repair

The comparison between the Force-Drift graphs of the undamaged specimen and the
retrofitted one (Figure 8.17) shows a good agreement between the two tests up to a drift
level of 3% (associated to the spalling of the repairing grout, Figure 8.18), which is
considered beyond the design drift level for precast concrete structures.

GS4U and GS4U-R Force-Drift Comparison
-100
-75
-50
-25
0
25
50
75
100
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
Drift (%)
L
a
t
e
r
a
l

f
o
r
c
e

(
k
N
)
GS4U
GS4U-R
T
h
e
o
r
e
t
i
c
a
l

y
i
e
l
d
P-

Figure 8.17 GS4U GS4U-R Force-Drift comparison







DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
116

Drift 1.0 % Drift 2.5% Drift 3.0%

Figure 8.18 GS4U-R drift pictures.

8.2 DDBD application
Once the moment - curvature relationships, which describe the different column to
foundation connections behavior, have been defined, the next step is the definition of the
associated equivalent viscous damping. This could be done relating the hysteretic
damping obtained with the Jacobsen approach (Eqn. 3.21) to the equivalent damping
equations (Grant et al. 2004). To check the suitability of this statement the ratio between
the hysteretic damping equations relative to the two sets of Takeda models, fat and
narrow, available in Grant et al. (2004) is compared with the same ratio obtained with a
Jacobsen approach. Figure 8.19 shows the results of this comparison: the two ratios
cannot be related directly, in fact the Jacobsen approach ratio depends on the ductility
reached while the Grant et al. (2004) ratio seems less dependent from the ductility level
but depends on the effective period.





PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION
117
Jacobsen and Grant damping ratio for Takeda hysteretic models:
"narrow" Takeda (=0.5;=0;r=0.05) and "fat" Takeda (=0.3;=0.6;r=0.05)
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1 2 3 4 5 6
Ductility ()
R
a
t
i
o
Hysteretic area Ratio
Grant Ratio - Teff=0.5
Grant Ratio - Teff=4.0

Figure 8.19 Hysteretic area and equivalent viscous damping ratio.

Applying the DDBD procedure without considering the actual equivalent viscous damping
values leads to the following results (Figure 8.20), which show how the DDBD procedure
does not predict correctly the structural behavior.

0 1 2 3 4 5 6 7
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
5
Roof drift Maximum Drift
R
o
o
f

d
r
i
f
t

(
%
)
Column size 1=60, 2=70, 3=80, 4=90, 5=100, 6=110 cm


2.5% drift
CS
PF
GS4
GS4B
GS4U
0 1 2 3 4 5 6 7
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
5
Roof drift Mean Drift
R
o
o
f

d
r
i
f
t

(
%
)
Column size 1=60, 2=70, 3=80, 4=90, 5=100, 6=110 cm


2.5% drift
CS
PF
GS4
GS4B
GS4U

Figure 8.20 NLTH analyses mean and maximum roof drift.






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
118
Other than the influence of the equivalent damping adopted, this behavior is due to the
way the structural yield displacement is computed. The yield curvature estimation
(Priestley, 2003) in Eqn. 3.7 has been derived from the moment-curvature analyses of
square columns with flexural reinforcement evenly distributed along the perimeter, with a
cross section size of 160 cm and with a cover to the flexural reinforcement of 5 cm. In the
experimental tests just shown the column dimension was 40 cm and the cover to the
flexural reinforcement coming from the foundation 8 cm. This latter case presents a ratio
between the effective depth and the column cross section size significantly less than the
former case, where the mentioned ratio is close to unity.

8.3 Yield curvature equation
To take into account a more effective yield curvature evaluation it is proposed to
substitute the effective depth (d) instead of the column cross section size (B) in the
aforementioned formula. Therefore the new yield curvature adopted is:
y
y
d

= (8.1)
Where
1
takes into account the effects of different parameters into the formula. To define

1
a series of moment-curvature analyses has been carried out varying the following
parameters for axial load ratio values
( )
2
/
ck
N f B = ranging from 0.05 to 0.3 (which is
the axial load ratio range of the structures considered in this research):

l
Longitudinal steel ratio
B Cross section size (square cross section)
c
cover
Concrete cover
f
ck
Concrete cubic resistance
f
y
Steel yield stress
f
u
/f
y
Steel overstrength ratio

The effect of three different longitudinal re-bars configurations has been taken into
account (Figure 8.21) with increasing number of re-bars equally distributed along the
cross section sides.





PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION
119

Case 1 4 Re-bars Case 2 8 Re-bars Case 3 12 Re-bars



Figure 8.21 Longitudinal re-bars distribution.

The moment curvature analyses have been carried out starting from the set of codes
Cumbia (Montejo et al. 2007). The assumptions made for the analyses are:

1. Plain sections remain plane (Navier Bernoulli hypothesis).
2. Perfect bond between concrete and steel.
3. Concrete tension strength is ignored in the analysis.
4. Concrete non-linear relationship according to the model proposed by Mander et
al. (1988) for unconfined and confined concrete.
5. Steel non-linear relationship according to King et al. (1986)
6. The axial force is applied at the section centroid

The yield curvature obtained from the analysis is the yield curvature of the bilinear
idealization (Figure 8.22) of the moment-curvature response, which is generally of
sufficient accuracy for design purposes.
The first branch is associated to the secant stiffness from the origin through first yield,
defined as the moment-curvature point (M
y
,
y
) when the extreme tension reinforcement
reaches the yield strain or when the extreme compression fiber reaches a strain of 0.002,
whichever occurs first. This line is extended until reaching a moment equal to the nominal
moment capacity (M
N
) defined as the moment when the extreme compression fiber strain
is 0.004 or when the extreme tension reinforcement strain is 0.015, whichever occurs
first. The nominal yield curvature is therefore obtained as
y
=
y
*M
N
/M
y
.
The second branch is obtained connecting the nominal yield point to the ultimate yield
point.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
120
'y u
My
MN
Mu
y
(Mu, ) u
(MN, ) y

Figure 8.22 Moment-curvature bilinear idealization.

In the following figures (Figure 8.23; Figure 8.24; Figure 8.25) the results obtained from
the parametric study are shown for the different longitudinal re-bars distribution.
The main parameters affecting the yield curvature are the axial load ratio and the
longitudinal steel ratio. The first one determines the slope of the curve while the second
rules the vertical offset. Considering the mean values of all data obtained, the yield
curvature can be expressed as function of the axial load ratio, which leads, through a
least square procedure, to the relationships shown in Figure 8.26.
The curve flattens (i.e. it gets independent from axial load ratio) as the number of re-bars
increases. This is in accord to the results found in the literature, where the relationship is
more or less constant due to the high size of the cross section and the consequent high
number of re-bars close spaced along the sides.





PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION
121

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on
l
Axial load ratio

f
ck
=40MPa, B=450mm, cover=55mm, =0.04,
f
y
=500MPa, f
u
/f
y
=1.3,
us
=0.07
(

y

d
)
/
(

y
)

l
=0.005

l
=0.01

l
=0.02

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on cross section size
Axial load ratio

f
ck
=40MPa,
l
=0.02, cover=55mm, =0.04,
f
y
=500MPa, f
u
/f
y
=1.3,
us
=0.07
(

y

d
)
/
(

y
)


B=300mm
B=450mm
B=600mm

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on concrete cover
Axial load ratio

f
ck
=40MPa,
l
=0.02, B=450mm, =0.04,
f
y
=500MPa, f
u
/f
y
=1.3,
us
=0.07
(

y

d
)
/
(

y
)


c=30mm
c=55mm
c=80mm

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on f
ck
Axial load ratio

l
=0.02, B=450mm, cover=55mm, =0.04,
f
y
=500MPa, f
u
/f
y
=1.3,
us
=0.07
(

y

d
)
/
(

y
)


f
ck
=30MPa
f
ck
=40MPa
f
ck
=50MPa

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on f
y
Axial load ratio

f
ck
=40MPa, B=450mm, cover=55mm,
l
=0.02,
f
u
/f
y
=1.3,
us
=0.07, =0.04
(

y

d
)
/
(

y
)


f
y
=450 MPa
f
y
=500 MPa
f
y
=550 MPa

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on f
u
/f
y
Axial load ratio

f
ck
=40MPa, B=450mm, cover=55mm,
l
=0.02,
f
y
=500MPa,
us
=0.07, =0.04
(

y

d
)
/
(

y
)


f
u
/f
y
=1.1
f
u
/f
y
=1.3
f
u
/f
y
=1.5

Figure 8.23 Case 1 parametric study results.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
122

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on
l
Axial load ratio

f
ck
=40MPa, B=450mm, cover=55mm, =0.04,
f
y
=500MPa, f
u
/f
y
=1.3,
us
=0.07
(

y

d
)
/
(

y
)

l
=0.01

l
=0.02

l
=0.03

l
=0.04

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on cross section size
Axial load ratio

f
ck
=40MPa,
l
=0.02, cover=55mm, =0.04,
f
y
=500MPa, f
u
/f
y
=1.3,
us
=0.07
(

y

d
)
/
(

y
)


B=300mm
B=450mm
B=600mm

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on concrete cover
Axial load ratio

f
ck
=40MPa,
l
=0.02, B=450mm, =0.04,
f
y
=500MPa, f
u
/f
y
=1.3,
us
=0.07
(

y

d
)
/
(

y
)


c=30mm
c=55mm
c=80mm

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on f
ck
Axial load ratio

l
=0.02, B=450mm, cover=55mm, =0.04,
f
y
=500MPa, f
u
/f
y
=1.3,
us
=0.07
(

y

d
)
/
(

y
)


f
ck
=30MPa
f
ck
=40MPa
f
ck
=50MPa

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on f
y
Axial load ratio

f
ck
=40MPa, B=450mm, cover=55mm,
l
=0.02,
f
u
/f
y
=1.3,
us
=0.07, =0.04
(

y

d
)
/
(

y
)


f
y
=450 MPa
f
y
=500 MPa
f
y
=550 MPa

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on f
u
/f
y
Axial load ratio

f
ck
=40MPa, B=450mm, cover=55mm,
l
=0.02,
f
y
=500MPa,
us
=0.07, =0.04
(

y

d
)
/
(

y
)


f
u
/f
y
=1.1
f
u
/f
y
=1.3
f
u
/f
y
=1.5

Figure 8.24 Case 2 parametric study results.





PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION
123

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on
l
Axial load ratio

f
ck
=40MPa, B=450mm, cover=55mm, =0.04,
f
y
=500MPa, f
u
/f
y
=1.3,
us
=0.07
(

y

d
)
/
(

y
)

l
=0.01

l
=0.02

l
=0.03

l
=0.04

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on cross section size
Axial load ratio

f
ck
=40MPa,
l
=0.02, cover=55mm, =0.04,
f
y
=500MPa, f
u
/f
y
=1.3,
us
=0.07
(

y

d
)
/
(

y
)


B=300mm
B=450mm
B=600mm

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on concrete cover
Axial load ratio

f
ck
=40MPa,
l
=0.02, B=450mm, =0.04,
f
y
=500MPa, f
u
/f
y
=1.3,
us
=0.07
(

y

d
)
/
(

y
)


c=30mm
c=55mm
c=80mm

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on f
ck
Axial load ratio

l
=0.02, B=450mm, cover=55mm, =0.04,
f
y
=500MPa, f
u
/f
y
=1.3,
us
=0.07
(

y

d
)
/
(

y
)


f
ck
=30MPa
f
ck
=40MPa
f
ck
=50MPa

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on f
y
Axial load ratio

f
ck
=40MPa, B=450mm, cover=55mm,
l
=0.02,
f
u
/f
y
=1.3,
us
=0.07, =0.04
(

y

d
)
/
(

y
)


f
y
=450 MPa
f
y
=500 MPa
f
y
=550 MPa

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence on f
u
/f
y
Axial load ratio

f
ck
=40MPa, B=450mm, cover=55mm,
l
=0.02,
f
y
=500MPa,
us
=0.07, =0.04
(

y

d
)
/
(

y
)


f
u
/f
y
=1.1
f
u
/f
y
=1.3
f
u
/f
y
=1.5

Figure 8.25 Case 3 parametric study results.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
124

0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence
Axial load ratio
(

y

d
)
/
(

y
)


Mean value



Case 1 (4 re-bars):

1.77 1.59
y
y
d

= = +
(8.2)
0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence
Axial load ratio
(

y

d
)
/
(

y
)


Mean value



Case 2 (8 re-bars):

0.93 1.84
y
y
d

= = +
(8.3)
0 0.1 0.2 0.3
0
0.5
1
1.5
2
2.5
(
y
d)/(
y
) dependence
Axial load ratio
(

y

d
)
/
(

y
)


Mean value



Case 3 (12 re-bars):

0.91 1.86
y
y
d

= = +
(8.4)
Figure 8.26 Mean value results.





PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION
125
8.4 Equivalent viscous damping equation re-calibration and results
To recalibrate the equivalent viscous damping equation, a new procedure which allows a
straightforward evaluation of the hysteretic damping has been developed. The new
procedure has been compared to the one available in the literature to underline the
advantages. The procedure adopted to calibrate the hysteretic damping expression given
by Grant et al. (2004) can be summarized by:

1. Select the hysteresis model.
2. Arbitrary assign mass (M) and yield displacement (
y
).
3. Choose the effective period (T
eff
) and the displacement ductility (

).
4. Calculate maximum target displacement (
max
), secant stiffness (K
eff
), maximum
and yield force (F
max
, F
y
) and initial stiffness (K
0
):

max y

= ;
2
2
4
eff
eff
M
K
T

=
; (8.5), (8.6)

max max eff
F K = ;
( )
max
1 1
y
F
F
r

=
+
;
0
y
y
F
K =

. (8.7), (8.8), (8.9)


Where r is the ratio between the post yield and elastic stiffness.
5. Determine an initial guess for the hysteretic damping based on the equations
available in the literature (Blandon, 2004).
6. For each ground motion record determine the elastic spectral displacement
S
D
(T
eff
,
eq
).
7. Iterate with time history analyses and a combination of regula falsi and bisection
method to determine the ground motion scaling factor (the first trial is
max
/S
D
) to
set the maximum displacement obtained from the non linear analyses to
max
.
8. Iterate with a combination of regula falsi and bisection method to determine
eq
so
that the equivalent linear elastic system with K
eff
and
eq
leads to a maximum
displacement equal to
max
for the ground motions as scaled before.
9. Repeat the procedure for different T
eff
and

.
10. Use a least square procedure to calibrate the hysteretic damping equation
parameters.
The procedure scheme is shown in the following flow chart (Figure 8.27).





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
126
Get
max, Keff , Fmax, Fy, K0
Ground motions
Non Linear Time History
Sdof non-linear System
M, Fy, K0
First trial
eq
Elastic Spectral Displacement
SD(Teff, eq)
= max / SD(Teff , eq)
x Ground motions
max_NLTH = max Change
Linear Time History
Sdof linear System
M, Keff, eq
max_LTH = max
Change
eq
Keep
eq
x Ground motions
Yes
No
Yes
No
START
END
Assign
M, y
Select
Teff ,

Figure 8.27 Flow chart: equivalent viscous damping calibration procedure (Grant et al. 2004)





PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION
127
The procedure proposed in this study, which in this chapter has been applied to the
Takeda hysteretic model and in the next chapter will be extended to others, is based on
the analysis of the force displacement inelastic response of single degree of freedom
systems and can be summarized by:

1. Choose the Force-displacement Takeda hysteretic model parameters whose
hysteretic damping relationship will be calibrated (in this case the parameters
based on the experimental tests finite element modeling).
2. Select the displacement ductility (

).
3. Obtain the elastic spectral displacement (
( )
_ D el
S T
) and the constant ductility
inelastic spectral displacement (
( )
_
,
D in
S T

) using the Inspect algorithm


(Carr 2006).
The inelastic spectrum (
( )
_
,
D in
S T

) refers to the SDOF systems elastic period,


while the hysteretic damping equation refers to the effective period.
The elastic and effective period relationship is:

0
1 ( 1)
eff
r
T T

+
=
(8.10)
4. Choose an effective period T
eff
. To obtain the equivalent viscous damping
associated to T
eff
and

, S
D_el
(T
eff
) needs to be compared with S
D_in
(T
0
,

) which
represents the inelastic spectral displacement of a SDOF inelastic system with
effective period T
eff
.
5. Adopt an appropriate equation to describe the elastic spectra dependence on the
damping value, for the ground motions considered in this research:

_
_
( , )
7.8
( , 5%) 2.8
D el
D el
S T
S T


=
= +
(8.11)
this relationship relates directly S
D_el
(T
eff
) to S
D_in
(T
0
,

) by means of the
equivalent viscous damping:

_ 0
_
( , )
7.8
( ) 2.8
D in
D el eff eq
S T
S T

=
+
(8.12)
Therefore, considering a viscous damping contribution of 5%, the hysteretic
damping is:





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
128
( ) ( )
2
2
_
_ 0 0
( ) ( )
, , 5 7.8 2.8 5 7.8 1
( ) ( )
D el eff D eff
hyst eff eq eff
D in D
S T S T
T T
S T S T


| |
(
| |
| |
|
( = = = |
|
|
|
(
\
\

\
(8.13)
6. The equivalent viscous damping equation parameters are obtained by means of
least square regression using the downhill simplex algorithm (Nelder &
Mead, 1965), based on the average value of the ground motion inelastic spectra.
7. Repeat for different T
eff
.
8. Repeat for different

.

The procedure scheme is shown in the following flow chart (Figure 8.28).

Select

START
Ground motions
Elastic Spectral Displacement SD_el(T)
Ineastic Spectral Displacement SD_in(T,)
Select
Teff
T0 = Teff 1+ r (-1)

hy (Teff, ) = 7.8
SD_in(T0)
2
SD_el(Teff )
-1
END

Figure 8.28 Flow chart: new equivalent viscous damping calibration procedure.





PHASE 4: DDBD AND FOUNDATION TO COLUMN CONNECTION
129
The results of the hysteretic damping calibration of the Takeda models, which better
describe the moment-curvature relationship of the test specimens, are summarized in
Table 8.8 where a,b,c and d are the equivalent viscous damping equation parameters as
defined in Grant et al. (2004).
Table 8.8. New hysteretic damping equation parameters
a b c d
CS 0.511 0.143 0.677 0.692
PF 0.432 0.180 0.705 0.685
GS4 0.350 0.226 0.719 0.681
GS4B 0.333 0.238 0.721 0.680
GS4U 2.356 0.027 0.634 0.703

Figure 8.29 shows, for different effective periods, the new ductility - hysteretic damping
relationships and the comparison with the previous equation.

1 1.5 2 2.5 3 3.5 4
0
5
10
15
20
Equivalent Viscous Damping T
eff
= 0.5 s

H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)
Takeda "fat" (Grant 2004)
CS
PF
GS4
GS4B
GS4U
1 1.5 2 2.5 3 3.5 4
0
5
10
15
20
Equivalent Viscous Damping T
eff
= 1.0 s

H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)
Takeda "fat" (Grant 2004)
CS
PF
GS4
GS4B
GS4U

1 1.5 2 2.5 3 3.5 4
0
5
10
15
20
Equivalent Viscous Damping T
eff
= 2.0 s

H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)
Takeda "fat" (Grant 2004)
CS
PF
GS4
GS4B
GS4U
1 1.5 2 2.5 3 3.5 4
0
5
10
15
20
Equivalent Viscous Damping T
eff
= 4.0 s

H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)
Takeda "fat" (Grant 2004)
CS
PF
GS4
GS4B
GS4U

Figure 8.29 Effective period and displacement ductility hysteretic damping relationship.






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
130
Figure 8.30 shows the results of the DDBD procedure with the modified yield curvature
relationship and equivalent viscous damping equation parameters.

0 1 2 3 4 5 6 7
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
5
Roof drift Mean Drift
R
o
o
f

d
r
i
f
t

(
%
)
Column size 1=60, 2=70, 3=80, 4=90, 5=100, 6=110 cm


2.5% drift
CS
PF
GS4
GS4B
GS4U
0 1 2 3 4 5 6 7
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
5
Roof drift Maximum Drift
R
o
o
f

d
r
i
f
t

(
%
)
Column size 1=60, 2=70, 3=80, 4=90, 5=100, 6=110 cm


2.5% drift
CS
PF
GS4
GS4B
GS4U

Figure 8.30 NLTH analyses mean and maximum roof drift

A good agreement is found between the DDBD prediction and the NLTH analyses. In this
particular case the results are more affected by the modification of the yield curvature
equation than the equivalent viscous damping equation, in fact the displacement ductility
and the effective period of the structure under examination are about 1.2 and 1.1 s
respectively: these values lead to an hysteretic damping close to the one associated to
the Takeda fat model available in the literature as it is clear in Figure 8.29.

8.5 Concluding remarks
In this chapter the results of experimental tests to evaluate the flexural behavior of some
column to foundation connections typical of the precast industry have been shown. The
test results have been used to determine the hysteretic model parameters to adopt in the
equivalent viscous damping equation re-calibration. A new equation describing the
column yield curvature has been proposed to overcome the drawbacks of the equation
available in the literature especially when applied to some types of foundation to column
connections. A new calibration procedure has been adopted to evaluate the equivalent
viscous damping equation parameters related to the column to foundation connections
analyzed in the experimental campaign. The DDBD procedure with the new yield
curvature and equivalent viscous damping equation has been applied with good results.


131
9. HYSTERTIC DAMPING EQUATION CALIBRATION


The equivalent viscous damping equation procedure presented in the previous chapter is
here extended to other hysteretic rules Takeda, Bilinear, Ramberg - Osgood and Flag
Shape hysteretic models (as in Grant et al. 2004), whose schematic representation is
shown in Figure 9.1. A modified version of the equivalent viscous damping equation is
proposed and the influence of the hysteretic parameters, the ground motion types and the
displacement response spectrum shape is evaluated.

F
1
=
k0
1 +
F
Fy
i
Fi
F-Fi
1
i =
k0
1 +
F-Fi
2 Fy
y u
Fy
k0
rk0
rlow k0
k0
F0
Fy -F0
Fy
= =
r - rlow
r - r x rlow
y u
Fy
F
k0
rk0
p
p
ku=k0

res y u
Fy
F
k0
rk0
p

Takeda Bilinear
Flag Shape Ramberg-Osgood
F


Figure 9.1 Hysteretic models relationship and parameters.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
132
9.1 Hysteretic damping calibration results
Before evaluating the hysteretic damping equations, the calibration procedure for the
Ramberg-Osgood hysteretic rule needs to be modified to take into account the
relationship between the elastic and effective period which differs from the one stated
before (Eqn. 8.10). The new relationship is found with the following steps:
1. From the monotonic loading relationship

1
0
1
i i
i
y
F F
K F

(
| |
(
= + |
|
(
\

, (9.1)
substituting K
0
with F
y
/
y
and

=
u
/
y
leads to:

1
1
u u u
y y y
F F
F F

(
| |

(
= = + |
|
(
\

(9.2)
2. Solve the previous equation and get A=F
u
/F
y
.
3. The effective stiffness is:

0
u i
eff
u y
F A F A
K K

= = =

(9.3)
4. The T
0
-T
eff
relationship is:

0 eff
A
T T

= (9.4)
A set of six statistically independent artificial records obtained using the Simqke algorithm
(Carr 2006) have been selected. Artificial records have been chosen instead of scaled
ground motions to better control the coefficient of variation (CoV, defined as the ratio
between standard deviation and mean value) along the spectral displacement and
spectral acceleration period range. The spectral acceleration and displacement
comparison is shown in Figure 9.2. To compare the results obtained with the new
calibration with the ones obtained by Grant et al. (2004), the artificial ground motions
have been created to have a spectral displacement shape compatible with the one
adopted by Grant et al. (2004), i.e. a displacement response spectrum linear up to a
period of 4 s. This does not reflect the Eurocode 8 (2004) displacement response
spectrum which presents a constant displacement region at 2 s.





HYSTERTIC DAMPING EQUATION CALIBRATION
133
The influence of the spectral displacement shape on the equivalent viscous damping
equation parameters and on the equation itself will be evaluated later in the chapter.

Spectral Acceleration Comparison
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
0 1 2 3 4
Period (s)
A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Mean
EC8
CoV
Spectral Displacement Comparison
0.0
0.2
0.4
0.6
0.8
1.0
0 1 2 3 4
Period (s)
D
i
s
p
l
a
c
e
m
e
n
t

(
m
)
Mean
EC8

Figure 9.2 Artificial records acceleration and displacement spectra.

After applying the calibration procedure for Takeda hysteretic models with different sets of
parameters, the hysteretic damping-effective period data are not well predicted
(Figure 9.3) by the actual hysteretic equation especially for low effective period values.
Therefore in this study the following new equation is proposed to better fit the data:
( )
1
a 1
d
hy eff
b
T c

| |
= +
|
\
(9.5)
For each of Figure 9.1 hysteretic rules, Figure 9.3 to Figure 9.6 show the results of this
new equation calibration and the comparison with the results obtained by Grant et
al. 2004.
Eqn. 9.5 parameters have been calibrated for different sets of hysteretic model
parameters as show in Table 9.1 to Table 9.4.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
134
Takeda hysteretic model

0 1 2 3 4
0
5
10
15
20
25
30
Hysteretic Damping

=2
T
eff
(s)
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model
0 1 2 3 4
0
5
10
15
20
25
30
Hysteretic Damping

=4
T
eff
(s)
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model

0 1 2 3 4
0
5
10
15
20
25
30
Hysteretic Damping

=6
T
eff
(s)
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model
1 2 3 4 5 6
0
5
10
15
20
25
30
Hysteretic Damping T
eff
=1.0s
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model

1 2 3 4 5 6
0
5
10
15
20
25
30
Hysteretic Damping T
eff
=2.5s
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model
1 2 3 4 5 6
0
5
10
15
20
25
30
Hysteretic Damping T
eff
=4.0s
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model

Figure 9.3 Takeda hysteretic damping equation calibration results.





HYSTERTIC DAMPING EQUATION CALIBRATION
135
Table 9.1. Takeda hysteretic damping equation parameters
Hysteretic model parameters Hysteretic damping equation parameters
r a b c d
0.005 0.1 0 1.331 0.303 -0.729 0.151
0.005 0.1 0.3 0.841 0.268 -0.493 0.26
0.005 0.1 0.6 0.355 0.262 0.282 0.522
0.005 0.3 0 0.704 0.485 -0.635 0.172
0.005 0.3 0.3 0.822 0.333 -0.571 0.204
0.005 0.3 0.6 1.032 0.331 -0.643 0.173
0.005 0.5 0 0.277 0.377 0.105 0.355
0.005 0.5 0.3 1.333 0.253 -0.683 0.142
0.005 0.5 0.6 0.972 0.098 0.024 0.422
0.05 0.1 0 0.529 0.296 -0.262 0.21
0.05 0.1 0.3 0.837 0.153 -0.165 0.248
0.05 0.1 0.6 0.696 0.128 0.203 0.344
0.05 0.3 0 0.737 0.096 0.359 0.306
0.05 0.3 0.3 0.762 0.124 0.059 0.281
0.05 0.3 0.6 0.739 0.159 -0.121 0.265
0.05 0.5 0 0.551 0.082 0.889 0.393
0.05 0.5 0.3 0.455 0.169 0.208 0.292
0.05 0.5 0.6 0.578 0.081 0.925 0.414
0.1 0.1 0 0.714 0.102 0.279 0.248
0.1 0.1 0.3 0.655 0.062 1.288 0.394
0.1 0.1 0.6 0.478 0.149 0.39 0.274
0.1 0.3 0 0.27 0.067 3.648 0.581
0.1 0.3 0.3 0.779 0.082 0.331 0.248
0.1 0.3 0.6 0.483 0.123 0.52 0.262
0.1 0.5 0 0.686 0.077 0.304 0.199
0.1 0.5 0.3 0.522 0.089 0.539 0.231
0.1 0.5 0.6 0.557 0.105 0.316 0.18





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
136
Bilinear hysteretic model

0 1 2 3 4
0
5
10
15
20
25
30
Hysteretic Damping

=2
T
eff
(s)
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model
0 1 2 3 4
0
5
10
15
20
25
30
Hysteretic Damping

=4
T
eff
(s)
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model

0 1 2 3 4
0
5
10
15
20
25
30
Hysteretic Damping

=6
T
eff
(s)
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model
1 2 3 4 5 6
0
5
10
15
20
25
30
Hysteretic Damping T
eff
=1.0s
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model

1 2 3 4 5 6
0
5
10
15
20
25
30
Hysteretic Damping T
eff
=2.5s
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model
1 2 3 4 5 6
0
5
10
15
20
25
30
Hysteretic Damping T
eff
=4.0s
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model

Figure 9.4 Bilinear hysteretic damping equation calibration results.





HYSTERTIC DAMPING EQUATION CALIBRATION
137
Table 9.2. Bilinear hysteretic damping equation parameters
Hysteretic model
parameter
Hysteretic damping equation parameters
r a b c d
0 0.272 0.097 2.842 0.526
0.05 1.59 0.075 0.175 -0.13
0.1 1.301 0.048 0.979 -0.092
0.15 0.452 0.12 1.053 -0.159
0.2 0.545 0.1 0.635 -0.023





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
138
Ramberg-Osgood hysteretic model

0 1 2 3 4
0
5
10
15
20
25
30
Hysteretic Damping

=2
T
eff
(s)
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model
0 1 2 3 4
0
5
10
15
20
25
30
Hysteretic Damping

=4
T
eff
(s)
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model

0 1 2 3 4
0
5
10
15
20
25
30
Hysteretic Damping

=6
T
eff
(s)
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model
1 2 3 4 5 6
0
5
10
15
20
25
30
Hysteretic Damping T
eff
=1.0s
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model

1 2 3 4 5 6
0
5
10
15
20
25
30
Hysteretic Damping T
eff
=2.5s
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model
1 2 3 4 5 6
0
5
10
15
20
25
30
Hysteretic Damping T
eff
=4.0s
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model

Figure 9.5 Ramberg-Osgood hysteretic damping equation calibration results.





HYSTERTIC DAMPING EQUATION CALIBRATION
139
Table 9.3. Ramberg-Osgood hysteretic damping equation parameters
Hysteretic model
parameter
Hysteretic damping equation parameters
a b c d
2 0.043 2.729 1.004 -0.841
3 0.241 1.339 -0.628 -0.103
4 0.215 4.563 -0.733 -0.064
5 0.231 1.546 -0.764 -0.071
7 0.206 43.832 -0.858 -0.062






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
140
Flag Shape hysteretic model

0 1 2 3 4
0
5
10
15
20
25
30
Hysteretic Damping

=2
T
eff
(s)
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model
0 1 2 3 4
0
5
10
15
20
25
30
Hysteretic Damping

=4
T
eff
(s)
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model

0 1 2 3 4
0
5
10
15
20
25
30
Hysteretic Damping

=6
T
eff
(s)
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model
1 2 3 4 5 6
0
5
10
15
20
25
30
Hysteretic Damping T
eff
=1.0s
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model

1 2 3 4 5 6
0
5
10
15
20
25
30
Hysteretic Damping T
eff
=2.5s
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model
1 2 3 4 5 6
0
5
10
15
20
25
30
Hysteretic Damping T
eff
=4.0s
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


TH Results
Best fit
Previous model

Figure 9.6 Flag Shape hysteretic damping equation calibration results.





HYSTERTIC DAMPING EQUATION CALIBRATION
141
Table 9.4. Flag shape hysteretic damping equation parameters
Hysteretic model parameter Hysteretic damping equation parameters
r a b c d
0.005 0.1 0.109 0.599 0.066 -0.38
0.005 0.3 0.13 0.73 0.147 -0.186
0.005 0.5 0.093 0.589 1.305 -0.076
0.05 0.1 0.054 0.378 1.178 -0.855
0.05 0.3 0.133 0.371 0.476 -0.335
0.05 0.5 0.141 0.269 1.31 -0.166
0.1 0.1 0.166 0.051 1.901 -1.17
0.1 0.3 0.213 0.095 1.275 -0.499
0.1 0.5 0.431 0.103 0.436 -0.137






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
142
9.2 Hysteretic model parameters influence on the damping value
In this paragraph the hysteretic model parameters influence on the hysteretic damping is
evaluated; the focus is on the Takeda hysteretic model but the results can be extended to
other hysteretic rules as well.
Figure 9.7 shows how the Takeda parameters (, and r) affect the hysteretic damping
value. The value does not affect the results for low effective period values and slightly
affects the response of SDOF systems with higher effective periods. The main
contribution to the hysteretic damping is given by both the and r values, with higher
dependence associated to higher effective period values.
Starting from this comparison, the hysteretic parameters influence for different types of
ground motions is sought, in particular for near and far field records.
Intuitively, for near field ground motions a greater influence of the r parameter is
expected, being near field records usually short in time and accompanied by a significant
velocity and acceleration pulse. This pulse is associated to the structural maximum
displacement response; if the structural response is mainly elastic before this pulse, the
inelastic behavior should be governed by the post yield stiffness parameter.
Far field ground motions have usually a higher duration and the records do not show the
near field pulse mentioned before, the structure goes into inelastic cycles before reaching
the maximum displacement response, therefore the inelastic response is more likely
controlled by all the hysteretic parameters and not only by the post yield stiffness.
To validate these statements, two natural ground motions have been selected. The
record data are summarized in Table 9.5 and ground acceleration, velocity and
displacement in Figure 9.8. Figure 9.9 shows the spectral displacement and spectral
acceleration records comparison.

Table 9.5. Near Field and Far Field ground motion data.
Record Location Station Date Magnitude Epicenter
Near Field Northridge Rinaldi 17/01/1994 6.7 5.54 km
Far Field Imperial Valley El Centro #8 15/10/1979 6.5 28.09 km






HYSTERTIC DAMPING EQUATION CALIBRATION
143

1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping r dependence (=0.3 ,=0.3, T
eff
=0.5s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


r=0.005
r=0.050
r=0.100
1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping r dependence (=0.3 ,=0.3, T
eff
=2.0s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


r=0.005
r=0.050
r=0.100

1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping dependence (r=0.05 ,=0.3, T
eff
=0.5s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


=0.1
=0.3
=0.5
1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping dependence (r=0.05 ,=0.3, T
eff
=2.0s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


=0.1
=0.3
=0.5

1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping dependence (=0.3 ,r=0.05, T
eff
=0.5s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


=0
=0.3
=0.6
1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping dependence (=0.3 ,r=0.05, T
eff
=2.0s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


=0
=0.3
=0.6

Figure 9.7 Takeda parameters influence on the hysteretic damping.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
144

Near Field Far Field
Acceleration
-1
-0.5
0
0.5
1
0 5 10 15
Time (s)
A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Acceleration
-1
-0.5
0
0.5
1
0 10 20 30 40
Time (s)
A
c
c
e
l
e
r
a
t
i
o
n

(
g
)

Velocity
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
0 5 10 15
Time (s)
V
e
l
o
c
i
t
y

(
m
/
s
)
Velocity
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
0 10 20 30 40
Time (s)
V
e
l
o
c
i
t
y

(
m
/
s
)
Displacement
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0 5 10 15
Time (s)
D
i
s
p
l
a
c
e
m
e
n
t

(
m
)
Displacement
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0 10 20 30 40
Time (s)
D
i
s
p
l
a
c
e
m
e
n
t

(
m
)
Figure 9.8 Near and far field acceleration, velocity and displacement comparison.





HYSTERTIC DAMPING EQUATION CALIBRATION
145

Near Field Far Field
Acceleration Spectrum
0
0.5
1
1.5
2
2.5
0 1 2 3 4
Period (s)
A
c
c
e
l
e
r
a
t
i
o
n

(
g
)

Acceleration Spectrum
0
0.5
1
1.5
2
2.5
0 1 2 3 4
Period (s)
A
c
c
e
l
e
r
a
t
i
o
n

(
g
)

Displacement Spectrum
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0 1 2 3 4
Period (s)
D
i
s
p
l
a
c
e
m
e
n
t

(
m
)

Displacement Spectrum
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0 1 2 3 4
Period (s)
D
i
s
p
l
a
c
e
m
e
n
t

(
m
)

Figure 9.9 Near and far field spectral acceleration and displacement comparison

Regarding the spectrum reduction factor (Eqn. 3.44), Priestely et al. (2007) proposed
the following equation suitable for near field records exhibiting the forward directivity
velocity pulse:
( )
0.25
( ) / (5%) 10/ 5
eq eq
SD SD
(
= = +

(9.6)
The factor has been calibrated for the two ground motion selected. This has been done
with a least square procedure applied in the period range 0 1.4 s and 1.4 4 s. The
results of the calibration are shown in Figure 9.10 and Figure 9.11.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
146
= SD(x)/SD(5%)
0.6
0.7
0.8
0.9
1
1.1
1.2
1.3
2 4 6 8 10 12 14 16
Equivalent viscous damping (%)
S
D
(
x
)
/
S
D
(
5
%
)
g = 10 (EC8)
g = 11.8 (Best fit)
Computed Data: mean 0-1.4 s
Computed Data: mean 1.4-4 s
0.5
( )
(5%) ( 5)
SD g
SD g

| |
=
|
+
\

Figure 9.10 Spectrum reduction factor for the far field record
= SD(x)/SD(5%)
0.6
0.7
0.8
0.9
1
1.1
1.2
1.3
2 4 6 8 10 12 14 16
Equivalent viscous damping (%)
S
D
(
x
)
/
S
D
(
5
%
)
g = 10 (Priestley et al. 2007)
g = 7.8 (Best fit)
Computed Data: mean 0-1.2 s
Computed Data: mean 1.2-4 s
0.25
( )
(5%) ( 5)
SD g
SD g

| |
=
|
+
\

Figure 9.11 Spectrum reduction factor for the near field record

The calibration of the hysteretic damping equation (Eqn. 9.5) for the Takeda hysteresis
rule have been carried out for the two records of Table 9.5. The results in terms of
hysteretic parameters dependence are shown in Figure 9.12 (near field) and Figure 9.13
(far field) where good correspondence to the previous statements is shown: the hysteretic
behavior of the near field record are ruled mainly by the r parameter, while in the far
field record, other than r, the influence of is significant.





HYSTERTIC DAMPING EQUATION CALIBRATION
147

1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping r dependence (=0.5 ,=0.6, T
eff
=1.0s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


r=0.005
r=0.100
1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping r dependence (=0.5 ,=0.6, T
eff
=2.0s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


r=0.005
r=0.100

1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping dependence (r=0.005 ,=0.6, T
eff
=1.0s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


=0.1
=0.5
1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping dependence (r=0.005 ,=0.6, T
eff
=2.0s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


=0.1
=0.5

1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping dependence (=0.5 ,r=0.005, T
eff
=1.0s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


=0
=0.6
1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping dependence (=0.5 ,r=0.005, T
eff
=2.0s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


=0
=0.6

Figure 9.12 Takeda parameters influence on the hysteretic damping for Near field record





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
148

1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping r dependence (=0.5 ,=0.6, T
eff
=1.0s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


r=0.005
r=0.100
1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping r dependence (=0.5 ,=0.6, T
eff
=2.0s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


r=0.005
r=0.100

1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping dependence (r=0.005 ,=0.6, T
eff
=1.0s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


=0.1
=0.5
1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping dependence (r=0.005 ,=0.6, T
eff
=2.0s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


=0.1
=0.5

1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping dependence (=0.5 ,r=0.005, T
eff
=1.0s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


=0
=0.6
1 2 3 4 5 6
0
5
10
15
20
25
Hysteretic Damping dependence (=0.5 ,r=0.005, T
eff
=2.0s)
Ductility
H
y
s
t
e
r
e
t
i
c

d
a
m
p
i
n
g

(
%
)


=0
=0.6

Figure 9.13 Takeda parameters influence on the hysteretic damping for Far field record





HYSTERTIC DAMPING EQUATION CALIBRATION
149
Other aspects related to the hysteretic damping associated to these records is the
significant damping decrease with the effective period increase. This behavior could be
related to the spectral displacement shape (Figure 9.9) which does not present the linear
monotonic increase of the records adopted in the previous chapter. Therefore the spectral
displacement shape could be a possible parameter to take into account in the hysteretic
damping evaluation.
To better understand how the shape of the spectral displacement affects the hysteretic
damping, another ground motion have been selected (Table 9.6) whose spectral
displacement shows a decreasing values region after an increasing one (Figure 9.14).
Table 9.6. Ground motion data.
Location Station Date Magnitude Epicenter
Livermore Mortan Territory Park 27/01/1980 5.5 8.1 km

Figure 9.14 shows the elastic displacement spectrum and the inelastic spectrum
associated to the Takeda hysteretic model (=0.3; =0.6; r=0.05) with a constant
displacement ductility value of 5.

Displacement spectrum
0.00
0.01
0.02
0.03
0.04
0 1 2 3 4
Elastic Period (s)
D
i
s
p
l
a
c
e
m
e
n
t

(
m
)
Elastic
Ductility 5

Figure 9.14 Elastic and inelastic spectral displacement.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
150
Looking at the calibration procedure stated in Chapter 8 (Figure 8.28) the abscissa values
of the inelastic spectrum need to be corrected according to Eqn. 8.10 in order to
determine the hysteretic damping. This correction leads to Figure 9.15.

Displacement spectrum
0.00
0.01
0.02
0.03
0.04
0 0.5 1 1.5 2 2.5 3 3.5 4
Effective Period (s)
D
i
s
p
l
a
c
e
m
e
n
t

(
m
)
Elastic
Ductility 5 (Time corrected)

Figure 9.15 Elastic and inelastic spectral displacement time corrected.

The hysteretic damping depends directly (Eqn. 8.13) from the square of the ratio between
the elastic spectral displacement and the inelastic (time corrected) one. Figure 9.15
shows how this ratio increases in the effective period range 0-1.1 s, decreases and
reaches 1 (which corresponds to an hysteretic damping equal to 0) in the effective period
range 1.1-1.65 s and is less than 1 for an effective period greater than 1.65 s.
In this last region the displacement of the elastic SDOF system with period T
eff
is lower
than the displacement of the inelastic SDOF system whose effective period is T
eff
.
Evaluating the hysteretic damping in such conditions leads to a negative value, which
means that the structure cannot reach the displacement ductility value supposed (5 in this
case) with an effective period T
eff
.
The actual ultimate displacement is found in two ways (Figure 9.16):
1. Maintain the displacement ductility value and decrease the effective period (i.e.
increase the effective stiffness).
2. Maintain the effective period (i.e. same effective stiffness) and reduce the
displacement ductility value.






HYSTERTIC DAMPING EQUATION CALIBRATION
151
y u
Fy
F y
inelastic
e
l
a
s
t
i
c

T
=
T
eff
y u
Fy
F y
e
l
a
s
t
i
c

T
=
T
eff
y u
Fy
F y
e
l
a
s
t
i
c

T
=
T
eff

Figure 9.16 Iterations to find the correct displacement ductility or the correct effective period.

1. 2.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
152
9.3 Concluding remarks
This chapter shows how the hysteretic damping depends on the type of ground motions
(near and far field) and on the shape of the displacement spectrum. Therefore the
hysteretic damping equation parameters need to be calibrated accordingly to the type of
seismic hazard associated to the site, selecting a significant set of ground motions.
In the case of displacement spectra with a constant and/or a decreasing value region,
other than a new parameters calibration the hysteretic damping equation suitability needs
to be checked, in particular the portion regarding the effective period dependence.
In the previous part of this research the hysteretic damping calibration has been carried
out for a set of ground motions whose displacement spectrum does not catch the
constant displacement region of the design spectrum adopted in Eurocode 8 (2004). For
the previous considerations it is correct to use the hysteretic damping equation, as it is,
up to an effective period correspondent to the corner period, for effective periods greater
than the corner period the hysteretic damping equation should be recalibrated based on
another set of ground motions, which catches the constant displacement region of the
design spectrum.

153
10. GROUND MOTION SCALING


In the DDBD validation and in the equivalent viscous damping calibration procedure it has
been noted how important is the choice of a suitable set of ground motions whose
characteristics are compatible with the earthquake scenario, whose mean displacement
spectrum well agree with the Code displacement spectrum adopted in the design
procedure and whose variability is controlled in the period range of interest in order to
limit the results variability, here measured by means of the Coefficient of Variation (CoV,
i.e. the ratio between the standard deviation and the mean value for each period or
frequency of the ground motion acceleration or displacement spectrum).
In this chapter, after a quick presentation of some of the record selection and scaling
methods to use for non linear time history analyses, a procedure is proposed to artificially
scale records, with the least amount of manipulation, in order to set the records spectral
shape CoV to a chosen constant value; this procedure is expected to reduce and control
the results variability arising from the set of ground motions chosen for the equivalent
viscous damping calibration procedure and for the non linear analyses of multi degree of
freedom systems.

10.1 Record selection and scaling
The most common way to choose ground motions for the non linear time history analyses
is to select, from available databases, real earthquake records based on magnitude and
distance and then scaling them to match the spectral acceleration evaluated at the
structural fundamental period, Sa(T
1
), obtained from an attenuation law and
corresponding to a given probability of exceedance. A possible drawback of this
procedure is that different bins of ground motions could lead to different non linear
responses: the spectral shape of records with the same Sa(T
1
) will affect the response of
multi degree of freedom and non-linear structures because the spectral values at other
periods will affect the response of higher modes and the non linear response when the
structural fundamental period lengthen. To overcome this drawback Baker & Cornell
(2005) proposed to add to the selection based on magnitude and distance another value
which takes into account the record spectral shape; this value is called and it is defined





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
154
as the number of standard deviations between the spectral acceleration of a recorded
ground motion at a given period and the spectral acceleration at the same period
computed via ground motion prediction models. Taking into account this value in the
records selection and scaling mentioned before leads to a reduction of the estimated
seismic demands scatter.
Shome et al. (1998) showed that to obtain an estimate of the median response

Z within a
certain factor X (i.e.

Z X Z ), with 95% confidence, the number of ground motion
needed is approximately
2 2
4 / n X = , where is the standard deviation of the natural
logarithms of the data. is mainly affected by the scaling method, therefore a scaling
method which leads to low values will require a smaller amount of ground motion
records. A possible way to reduce is to scale the response spectrum of each record in a
bin to the mean spectrum of the bin itself. This forced reduction of , although does not
affect the median value of the data, could lead to a sensibly different actual of the
response, especially when this number is required; like for example in the design criterion
which calls for the 84
th
percentile demand or in a probabilistic design.
Another way of selecting earthquake records that reduces variability in the results has
been proposed by Baker & Cornell (2006), who suggested to choose records with an
acceleration spectrum similar in shape to the conditional response spectrum which is
related to a given level of spectral acceleration at the structure fundamental period and its
associated magnitude, distance and obtained from probabilistic seismic hazard de-
aggregation (refer to Baker & Cornell 2006 for detailed insight).
Other types of records to adopt in the non linear analyses are the synthetic records,
which are obtained from seismological models. The synthetic record generation requires
good knowledge of the earthquake source, path and site parameters, which are not
always available and therefore have not been taken into account in this research.
Another possible way to decrease the CoV is to create artificial records statistically
independent with a Fourier decomposition of the target spectrum for the time invariant
frequency content and a time dependent modulating function to simulate the non
stationary earthquake characteristics. The drawback of this procedure is that the ground
motions obtained can be seen as a sum of harmonic functions but with a phase sequence
not associated to physical meaning. This procedure does not take into account the
temporal variation in the earthquake record frequency composition due to the arrivals of
different types of seismic waves at different time instants and to other phenomena. Other





GROUND MOTION SCALING
155
than that, non linear systems respond differently to ground motions with time invariant
frequency content even if there is no difference in the response of linear systems.
Other scaling methods are available in the literature; the one adopted in this research has
been proposed by Mukherjee & Gupta (2002). The authors applied a wavelet-based
procedure to decompose a recorded ground motion into a desired number of time
histories with non-overlapping frequency contents. After that each of the time history is
scaled in amplitude with an iterative process so that the acceleration spectrum of the time
history ensemble will match a specified target spectrum. In this way the temporal
variations of the accelerogram frequency content is retained in the synthesized one.
This method does not bypass the issues related to records selection, because it implicitly
assumes that the ground motion selected has been recorded under similar seismic
environment and site conditions of the earthquake scenario under interest and therefore
the seismic waves exhibit desired temporal characteristics.

10.2 Constant Variance Spectrum Matching procedure
Based on the work of Mukherjee & Gupta (2002), the procedure proposed hereon is to
adopt that method to scale the selected records in order to set the CoV of the
acceleration spectra ensemble to a chosen constant value: Constant Variance Spectrum
Matching procedure (CVSM). The idea of taking into account the record variability in the
ground motions bin has been recently exploited by Kottke & Rathje (2008), who proposed
a method to select and linear scale records in a bin in order to match both the median
value and the standard deviation according to attenuation laws. The records scaling has
been done in order to minimize the root mean square error between the target standard
deviation and the standard deviation of the records ensemble in the log space. The
method has been shown to work fairly well in the case the standard deviation record in
the log space is fairly constant, although the record selection, based on spectral shape
suitability, is not straightforward and does not take into account the earthquake
characteristics like magnitude, distance, topography, duration etc.
The method proposed here allows to scale records and match a predetermined CoV with
less restriction from a spectral shape point of view respect to the Kottke & Rathje (2008)
method. Before applying the scaling method, an appropriate selection of the records
should be made. According to the earthquake scenario under interest, the bin selected
should be composed of records, if they exist, that come ideally from a similar faulting,
subsoil type, magnitude and distance, and hopefully topography. The further the records





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
156
selected will be from the earthquake scenario under interest the more the scaling
procedure proposed will affect the records to match the desired scenario. Another aspect
that could be taken into account is the earthquake duration which can affect the nonlinear
analyses results.
The CVSM procedure proposed can be described as:
1. Define the Code target spectrum.
2. Choose the constant coefficient of variation (CoV) for the acceleration spectra
ensemble of the records selected; the CoV could be chosen for example as the
value resulting from attenuation laws at the first mode period of the structure.
3. Define the new target spectrum for each ground motion.
It has to be noted that this research adopted the Wavgen algorithm (Mukherjee &
Gupta, 2002) for the scaling procedure; this algorithm requires an input time
history and a target acceleration spectrum defined at 97 period points, ranging
from 0.044 to 11.313 s and equally spaced in a logarithmic scale (to assign equal
weight to short and long periods). For each of those relevant periods the records
acceleration spectra values are evaluated and sorted in ascending order to match
a normal distribution whose mean value is the Code acceleration spectrum and
the CoV is the one chosen. A lognormal distribution can be used as well, in that
case the same procedure applies considering the log value of the acceleration
response spectra to match a normal distribution whose mean value is the log of
the Code acceleration spectrum.
In this research the results of sets of 12 records are presented, although other
numbers of ground motions can be used (for example 20 records, which is seen
as a more suitable number in the case of statistical analyses).
Starting with a normal distribution with a unitary standard deviation and zero
mean, the acceleration spectra values are distributed in subsets of 1, 2, 3, 3, 2, 1
in the case of 12 records (1, 2, 3, 4, 4, 3, 2, 1 for 20 records) as in Figure 10.1.
The same value of the standard normal probability density function is assigned at
each record.





GROUND MOTION SCALING
157

0
0.1
0.2
0.3
0.4
0.5
-3 -2 -1 0 1 2 3
Mean = 0; STD = 1;
Total 12 Ground Motions
1 GM
1 GM
2 GM
3 GM 3 GM
2 GM
Ai
0
0.1
0.2
0.3
0.4
0.5
-3 -2 -1 0 1 2 3
Mean = 0; STD = 1;
Total 20 Ground Motions
1 GM
1 GM
2 GM
3 GM
4 GM 4 GM
3 GM
2 GM
Ai

Figure 10.1 Ground motions normal distribution for sets of 12 and 20 records.

The coefficients A
i
are the values to add or subtract to the Code acceleration
spectrum in order to obtain a set of new spectra with unitary standard deviation
and with mean value the Code spectrum. To assign a chosen coefficient of
variation (CoV) along the spectrum, the A
i
coefficients need to be multiplied by
the standard deviation (STD) associated to the chosen CoV. For the generic
ground motion GM
i
the target spectrum value at the period T is:
( ) ( ) ( ) ( )
( )( )
i
(GM ) (Code) (Code) (Code)
a target a i a a i
(Code)
a i
S =S +STD A =S +CoV S A =
=S 1 CoV A
T T T T
T

+
(10.1)
The A
i
values are 1.6 (1+1 ground motions), 1.09 (2+2 ground motions) and
0.6 (3+3 ground motions) in the case of 12 records set and 1.74 (1+1 ground
motions), 1.29 (2+2 ground motions), 0.92 (3+3 ground motions), 0.53
(4+4 ground motions) for the 20 records set.
4. Apply the Wavgen algorithm (Mukherjee & Gupta, 2002) to the original time
histories in order to match the new target acceleration spectra.

The procedure proposed is seen as a possible way of controlling the results variability
and reducing the results dependence from the set of ground motions chosen and the
scaling method adopted. Controlling the dispersion of the input ground motions set with
this methodology could be suitable for the equivalent viscous damping equation
calibration procedure, for the design validation of multi degree of freedom systems and





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
158
for the probabilistic hazard assessment (like drift assessment) where accurate estimates
of both the response mean and dispersion are needed.
Another feature that it has been tried to add to the CVSM procedure is to minimize the
ground motion scaling in order to keep the scaled record acceleration spectral shape as
close as possible to the original one. It has to be noted that CVSM does not constitute a
ground motion selection criterion but it applies right after it, because it implies that the
earthquake records waves exhibit desired temporal characteristics, although the
procedure can widen the range of suitable records because for example it can adjust and
take into account ground motion recorded in slightly different site conditions.

10.3 CVSM application
The procedure proposed has been applied to two bins of 12 ground motions, taken from
a previous selection adopted in the preliminary design of the structure shown in the
previous chapter. The records selection and the CoV (25%) have been chosen to show
how the method works. In a real application it will be more appropriate to seek for typical
features that apply to the site under interest as mentioned before.
The records have been selected to match, after scaling, the 5% damping acceleration
spectrum (ASCE 7-05) for the Design Basis Earthquake with a probability of exceedance
of 10% in 50 years of a site located in Berkeley (CA USA) with a soil type C. Both
orthogonal horizontal components of the earthquake records selected have been
considered as single records to reduce by half the number of records required. This is
suitable for the analyses of single degree of freedom systems or multi degree of freedom
planar systems. In the case of applications requiring three orthogonal earthquake
components it is possible to apply the CVSM procedure to each component or to apply it
to one single component and to adopt the scale factors obtained to scale the other
components. The records selected are shown in Table 10.1and Table 10.2.





GROUND MOTION SCALING
159

Table 10.1. Earthquake records Bin 1.
Code Earthquake M Epicenter
GM 1 - 2 1987 Superstitn Hills - Parachute Test Site 6.7 16.0 km
GM 3 - 4 1968 Borrego Mountain - El Centro Array #8 6.8 70.8 km
GM 5 - 6 1992 Erzican (Turkey) - Erzican 6.9 9.0 km
GM 7 - 8 1941 Northen California - Ferndale City Hall 6.4 28.8 km
GM 9 - 10 1979 Imperial Valley - El Centro Array #8 6.5 28.1 km
GM 11 - 12 1994 Northridge - Rinaldi Receiving Station 6.7 5.5 km

Table 10.2. Earthquake records Bin 2.
Code Earthquake M Epicenter
GM 1 - 2 1986 Taiwan 31 Smart1 O07 7.3 39.0 km
GM 3 - 4 1991 Georgia 19 iri 6.2 36.4 km
GM 5 - 6 1989 Loma Prieta LGPC 6.9 18.5 km
GM 7 - 8 1992 Cape Mendocino Rio del Overpass 7.1 22.6 km
GM 9 - 10 1989 Loma Prieta Saratoga-Aloha Ave. 6.9 27.2 km
GM 11 - 12 1952 Kern County Taft Lincoln School 7.4 41.0 km

Before applying the CVSM procedure, each record has been scaled, multiplying the time
history by a factor determined with a least squares procedure, to match the Code target
spectrum in the period range 0.2 1.5 s here referred as Conventional scaling. This
scaling procedure is usually done in practice to take into account the higher modes
contribution and the structural fundamental period (here assumed 1 s) lengthening due to
the non linear structural behavior (ASCE 7-05). Figure 10.2 shows the results of the
scaling procedures.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
160

CVSM scaling procedure Conventional scaling procedure
B
i
n

1

10
1
10
0
10
1
10
2
10
1
10
0
10
1
Period (s)
A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Spectra Comparison


Elastic spectrum
GM after CVSM

10
2
10
1
10
0
10
1
10
2
10
1
10
0
10
1
Period (s)
A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Spectra Comparison


Elastic spectrum
GM after scaling

B
i
n

2

10
1
10
0
10
1
10
2
10
1
10
0
10
1
Period (s)
A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Spectra Comparison


Elastic spectrum
GM after CVSM

10
2
10
1
10
0
10
1
10
2
10
1
10
0
10
1
Period (s)
A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Spectra Comparison


Elastic spectrum
GM after scaling

Figure 10.2 Conventional and CVSM scaling for Bin 1 and Bin 2.

For each record of Bin 1, Figure 10.3 shows the acceleration spectra before and after
(continues and dashed lines respectively) the CVSM procedure application compared to
the Code spectrum (dotted line).






GROUND MOTION SCALING
161
10
0
0
0.5
1
1.5
2
2.5
GM 1
S
a

(
g
)
10
0
0
0.5
1
1.5
2
2.5
GM 2
10
0
0
0.5
1
1.5
2
2.5
GM 3
10
0
0
0.5
1
1.5
2
2.5
GM 4
S
a

(
g
)
10
0
0
0.5
1
1.5
2
2.5
GM 5
10
0
0
0.5
1
1.5
2
2.5
GM 6
10
0
0
0.5
1
1.5
2
2.5
GM 7
S
a

(
g
)
10
0
0
0.5
1
1.5
2
2.5
GM 8
10
0
0
0.5
1
1.5
2
2.5
GM 9
10
0
0
0.5
1
1.5
2
2.5
GM 10
Period (s)
S
a

(
g
)
10
0
0
0.5
1
1.5
2
2.5
GM 11
Period (s)
10
0
0
0.5
1
1.5
2
2.5
GM 12
Period (s)

Figure 10.3 Bin 1 ground motions before and after CVSM.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
162
For each record of Bin 1, Figure 10.4 shows the acceleration spectra scaling factors (SF)
values at each period for the CVSM and the Conventional scaling procedures (dashed
and dotted lines respectively). The results show the CVSM procedure effort to limit the SF
values along the periods (i.e. trying to maintain the original spectral shape), although
records GM 7 and GM 8 present high SF due to their lack of energy in the lower
frequency region.

0 2 4
0.5
1
1.5
2
GM 1
S
F
0 2 4
1
2
3
4
GM 2
0 2 4
2
4
6
8
GM 3
0 2 4
0
5
10
15
20
GM 4
S
F
0 2 4
0
1
2
3
GM 5
0 2 4
0.5
1
1.5
2
GM 6
0 2 4
0
10
20
30
GM 7
S
F
0 2 4
0
10
20
30
40
GM 8
0 2 4
1
1.5
2
2.5
3
GM 9
0 2 4
1
2
3
4
GM 10
Period (s)
S
F
0 2 4
0
0.5
1
1.5
GM 11
Period (s)
0 2 4
0.5
1
1.5
2
GM 12
Period (s)

Figure 10.4 Bin 1 scaling factors for CVSM and Conventional scaling procedures.





GROUND MOTION SCALING
163
Figure 10.5 shows the comparison between the ground motion GM 11 of set 1 (1994
Northridge - Rinaldi Receiving Station) before and after the application of CVSM.

0 5 10
0.5
0
0.5
Acceleration Comparison
A
c
c

(
g
)


Before scaling After scaling
1 2 3 4 5
0.5
0
0.5
A
c
c

(
g
)
Time (s)

0 5 10
0.5
0
0.5
1
1.5
Velocity Comparison
V
e
l

(
m
/
s
)


Before scaling After scaling
1 2 3 4 5
0.5
0
0.5
1
1.5
V
e
l

(
m
/
s
)
Time (s)

0 1 2 3 4
0
0.5
1
1.5
2
Spectral Acceleration Comparison
Period (s)
S
a

(
g
)


Before scaling
After scaling

0 1 2 3 4
0
0.5
1
1.5
2
Spectral Velocity Comparison
Period (s)
S
v

(
m
/
s
)


Before scaling
After scaling

0 5 10
0.2
0
0.2
0.4
0.6
0.8
1
1.2
1.4
Displacement Comparison
D
i
s
p

(
m
)
Time (s)


Before scaling After scaling

0 1 2 3 4
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
Spectral Displacement Comparison
Period (s)
S
d

(
m
)


Before scaling
After scaling

Figure 10.5 GM 11 comparison before and after CVSM application.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
164
The previous Figure 10.5 shows the success of the scaling procedure to retain the record
frequency content although a base line correction could be necessary as seen in the
displacement time history. Similar results have been obtained for Bin 2 and are not shown
for sake of simplicity.
The CVSM and Conventional scaled ground motions have been applied to a single
degree of freedom system governed by the Takeda hysteresis rule (refer to Chapter 5),
with the parameters =0, =0 and r=0.01, suitable to describe the non linear behavior of
reinforced concrete structures. In order to compare the results between the CVSM and
conventional scaling procedures, constant ductility spectra have been created with the
Inspect algorithm (Carr 2006) for displacement ductility (

) 1, 3 and 6 and with a


damping ratio of 5%.
The results in terms of mean values of the acceleration spectra, the displacement spectra
and the C values (defined as the ratio between the inelastic and elastic displacement at
each period) are shown in Figure 10.6 and Figure 10.7 for Bin 1 and Figure 10.8 and
Figure 10.9 for Bin 2. These Figures show that the constant CoV imposed to the input
ground motions ensembles by the CVSM procedure is overall well maintained in the
output results compared to the Conventional scaling procedure. Therefore the CVSM
procedure succeeds in controlling the results variability as required.





GROUND MOTION SCALING
165

CVSM scaling procedure Conventional scaling procedure
10
0
10
0
C

as function of

Time period (s)


C

= 1

= 3

= 6

10
0
10
0
C

as function of

Time period (s)


C

= 1

= 3

= 6

10
1
10
0
10
1
10
0
Spectral Acceleration
Period (s)
S
a

(
g
)

= 1

= 3

= 6

10
1
10
0
10
1
10
0
Spectral Acceleration
Period (s)
S
a

(
g
)

= 1

= 3

= 6

0 1 2 3 4
0
0.2
0.4
0.6
0.8
1
Spectral Acceleration CoV
Period (s)
C
o
V

= 1

= 3

= 6

0 1 2 3 4
0
0.2
0.4
0.6
0.8
1
Spectral Acceleration CoV
Period (s)
C
o
V

= 1

= 3

= 6

Figure 10.6 Bin 1: C and spectral acceleration comparison.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
166

CVSM scaling procedure Conventional scaling procedure
0 1 2 3 4
0
0.2
0.4
0.6
0.8
1
Spectral Displacement
Period (s)
S
D

(
m
)

= 1

= 3

= 6

0 1 2 3 4
0
0.2
0.4
0.6
0.8
1
Spectral Displacement
Period (s)
S
D

(
m
)

= 1

= 3

= 6

0 1 2 3 4
0
0.2
0.4
0.6
0.8
1
Spectral Displacement CoV
Period (s)
C
o
V

= 1

= 3

= 6

0 1 2 3 4
0
0.2
0.4
0.6
0.8
1
Spectral Displacement CoV
Period (s)
C
o
V

= 1

= 3

= 6

Figure 10.7 Bin 1: Spectral displacement comparison.





GROUND MOTION SCALING
167

CVSM scaling procedure Conventional scaling procedure
10
0
10
0
C

as function of

Time period (s)


C

= 1

= 3

= 6

10
0
10
0
C

as function of

Time period (s)


C

= 1

= 3

= 6

10
1
10
0
10
1
10
0
Spectral Acceleration
Period (s)
S
a

(
g
)

= 1

= 3

= 6

10
1
10
0
10
1
10
0
Spectral Acceleration
Period (s)
S
a

(
g
)

= 1

= 3

= 6

0 1 2 3 4
0
0.2
0.4
0.6
0.8
1
Spectral Acceleration CoV
Period (s)
C
o
V

= 1

= 3

= 6

0 1 2 3 4
0
0.2
0.4
0.6
0.8
1
Spectral Acceleration CoV
Period (s)
C
o
V

= 1

= 3

= 6

Figure 10.8 Bin 2: C and spectral acceleration comparison.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
168

CVSM scaling procedure Conventional scaling procedure
0 1 2 3 4
0
0.2
0.4
0.6
0.8
1
Spectral Displacement
Period (s)
S
D

(
m
)

= 1

= 3

= 6

0 1 2 3 4
0
0.2
0.4
0.6
0.8
1
Spectral Displacement
Period (s)
S
D

(
m
)

= 1

= 3

= 6

0 1 2 3 4
0
0.2
0.4
0.6
0.8
1
Spectral Displacement CoV
Period (s)
C
o
V

= 1

= 3

= 6

0 1 2 3 4
0
0.2
0.4
0.6
0.8
1
Spectral Displacement CoV
Period (s)
C
o
V

= 1

= 3

= 6

Figure 10.9 Bin 2: Spectral displacement comparison.

Figure 10.10 shows the comparison between the results of the two independent sets of
ground motions for the two scaling procedures. It is seen how the results obtained from
the Conventional scaling procedure are affected by the records selection with peak
differences up to 100% for the higher ductility values. The CVSM procedure succeeds in
reducing considerably the variability dictated by the record selection and leads to good
results for all the ductility values.





GROUND MOTION SCALING
169

CVSM scaling procedure Conventional scaling procedure
0 1 2 3 4
0
0.5
1
1.5
2
C

GMset dependence
Period (s)
C


s
e
t
1

/

C


s
e
t
2

= 1

= 3

= 6

0 1 2 3 4
0
0.5
1
1.5
2
C

GMset dependence
Period (s)
C


s
e
t
1

/

C


s
e
t
2

= 1

= 3

= 6

0 1 2 3 4
0
0.5
1
1.5
2
Spectral Acceleration GMset dependence
Period (s)
S
a

s
e
t
1

/

S
a

s
e
t
2

= 1

= 3

= 6

0 1 2 3 4
0
0.5
1
1.5
2
Spectral Acceleration GMset dependence
Period (s)
S
a

s
e
t
1

/

S
a

s
e
t
2

= 1

= 3

= 6

0 1 2 3 4
0
0.5
1
1.5
2
Spectral Displacement GMset dependence
Period (s)
S
D

s
e
t
1

/

S
D

s
e
t
2

= 1

= 3

= 6

0 1 2 3 4
0
0.5
1
1.5
2
Spectral Displacement GMset dependence
Period (s)
S
D

s
e
t
1

/

S
D

s
e
t
2

= 1

= 3

= 6

Figure 10.10 Set 1 and Set 2 - CVSM and Conventional scaling comparison.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
170
10.4 Concluding remarks
A procedure to reduce the variability of the non linear analyses output results due to the
choice of the input ground motions has been proposed. The procedure, Constant
Variance Spectrum Matching procedure, assigns a constant coefficient of variation to the
earthquake records ensemble and has shown good results when applied to two bins of
ground motions used in the analysis of a single degree of freedom non linear system,
although its suitability in reducing the results variability should be checked for multi
degree of freedom systems and for different hysteresis rules as well, evaluating the
variability in structural response for additional quantities like floor accelerations, peak
base shear and cumulative hysteretic energy among others.
The procedure has been adopted for the deterministic seismic analysis where the ground
motion selection described a single earthquake scenario. The CVSM could be suitable for
probabilistic analyses because it will be possible to select and scale different bins to
match the different earthquake scenarios whose the Uniform Hazard Spectrum (UHS) is
made of. Where the UHS is defined by having a spectral acceleration value at each
period with an exceedance probability equal to the specified target probability. UHS does
not represent a spectrum caused by a single earthquake at a single site because usually
the high frequency and the low frequency portions come from different events, small
nearby earthquakes for the former and larger and distant earthquake for the latter.
At last the procedure should be validated for near fault records as well, because these
kind of records display features like high-energy pulses that cannot be introduced by
linear scaling.

171
11. ROCKING WALLS IN PRECAST CONCRETE STRUCTURES


After the application of DDBD procedure to precast structures with classical lateral force
resisting systems (i.e. fixed end columns with or without the contribution of the top column
to beam connection), this chapter exploits the use of rocking walls as an alternative
resisting system to use in precast structures. Rocking walls have self centering properties
(given by post tensioning unbonded tendons) and accommodate the seismic lateral
displacement demand with a base rotation which leads to only one concentrated opening
at the foundation to wall joint compared to the crack spreading and damage typical of the
plastic region of classical reinforced concrete walls. This chapter deals with the problem
of the base sliding typical of these walls, with the definition of a moment rotation
relationship to use in the non linear analyses and with the revisiting and extension of the
equations of motion especially to determine the rocking period of the system whose
relation to the design procedure can be exploited as an extension of this research.

11.1 Rocking walls: an introduction
The first analytical study on the rocking motion of structures has been reported by
Housner (1963) to try to explain why during the Valdivia earthquake (Chile 22/05/1960)
several golf-ball-on-a-tee types of elevated water tanks survived the shaking despite the
appearance of instability, while much more stable-appearing reinforced-concrete elevated
water tanks were severely damaged. This apparently anomalous behavior has been
explained studying the dynamics of a rigid block resting upon a rigid horizontal base and
excited into rocking motion. The dynamic characteristics of these types of structures are
sensibly different from non-linear elastic structures. The author showed the low amount of
shear force and bending moment generated during an earthquake designing slender
structures to act as rigid blocks, however it was not clear how to reduce the probability of
overturning through the design procedure. The advantage of rocking in a structure is the
self-centering upon unloading and lack of residual drifts after an earthquake. A typical
example reported as design of a structure in controlled rocking conditions is the rail
bridge over the South Rangitikei River in New Zealand, where the 70 m tall piers have
been designed to allow a controlled base uplift and rocking in the transverse direction.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
172
Rocking solutions are usually applied to precast concrete systems although recent efforts
have been made to integrate the rocking self-centering concept to steel moment resisting
frames (Garlock et al. 2007).
To increase the self centering capacity, rocking systems have been coupled with
unbonded post tensioned tendons which provide restoring force with considerably less
concrete tension cracking in the system, if compared to monolithic solutions, as the
concrete is not bonded to the tendons and does not go into tension. These systems
perform well under the self centering point of view but low energy dissipation is
associated to the rocking process, mainly related to radiation damping and concrete
crushing in the impact region. The low energy dissipation could lead to greater system
displacements and to a higher number of high displacement peaks if compared to
monolithic solutions (Kurama 2002). To increase the energy dissipation without
jeopardize the self centering capacity, the systems can be coupled with frictional dampers
(Priestley et al. 1999), viscous dampers (Kurama 2000, Marriott et al. 2008) and mild
steel dissipation devices (Kurama 2002, Holden et al. 2003, Restrepo & Rahman. 2007,
Marriott et al. 2008). When such additional energy dissipation devices are placed, the
system is called hybrid.

11.2 Rocking walls experimental tests in the literature
The self-centering idea has been successfully applied and tested in the US-PRESSS
Program (PREcast Seismic Structural System, Priestley et al. 1999) as an evolution of
the rocking concepts developed by Priestley & Tao (1993) and Stone et al. (1995). In that
project a 60% scale five-story building prototype was constructed and tested under
pseudo-dynamic conditions to investigate the behavior of moment resisting frames
prestressed with partially unbonded tendons and mild steel to provide energy dissipation.
In the same program adjacent precast rocking walls with frictional dampers, placed
between the two adjacent walls, and post-tensioned unbonded tendons connecting the
walls top to the foundation have been tested.
The topic of this research is the study of precast concrete hybrid wall lateral force
resisting system as an alternative to conventional cast in place constructions.
Experimental tests to better understand the behavior of these systems are reported in the
literature. Other than the pseudo-dynamics tests made in the US-PRESSS Program,
monotonic and cyclic quasi static tests have been carried out to evaluate the base shear
top displacement relationship applying a lateral load, by means of an actuator, at the





ROCKING WALLS IN PRECAST CONCRETE STRUCTURES
173
point corresponding to the resultant of a lateral force distribution according to the flexural
first mode of vibration. Among these tests Kurama et al. (1999) tested unbonded post-
tensioned precast concrete walls, made by the superposition of one story precast panels.
The toe regions were heavily confined by means of spiral reinforcements and stirrups.
Holden et al. (2003) added milled mild steel re-bars across the base joint to provide a
satisfactory level of hysteretic damping. The authors detailed the wall reinforcement
layout according to the internal force flow obtained from a strut and tie model analysis
and adopted fiber reinforced concrete for the test unit. The same type of energy
dissipation bars have been used to test hybrid walls by Restrepo & Rahman (2007) with a
wall reinforcement layout made by low amount of steel ratio in the horizontal ( = 0.0025)
and vertical ( = 0.0084) directions and stirrups confined region at the wall toes. The
results outlined the main states of behavior under lateral loads and the flag shape
hysteresis loop typical of these hybrid systems (Figure 11.1).

Displacement
1
2
3
4 5
6
L
o
a
d

Figure 11.1 Rocking wall behavior states

According to Figure 11.1 the main behavior states can be summarized in:
1. Decompression, which identifies the initiation of gap opening at the wall base to
foundation joint.
2. Softening or geometric non-linearity, which is associated to the beginning of
significant reduction of the wall lateral stiffness due to gap opening or nonlinear
behavior of the toe concrete in compression (depending on the initial level of
vertical load due to gravity and post tensioning). Marriott et al. (2008) identify this
state with the neutral axis at mid depth of the wall section.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
174
3. Yielding of mild steel reinforcement, with consequent further decrease of the wall
lateral stiffness.
4. Yielding of post tensioning reinforcement. The post tensioning steel reaches the
limit of proportionality, the wall self centering property can be reduced.
5. Rupture of mild steel reinforcement, which can be avoided by controlling the steel
strain by means of an unbonded length.
6. Failure state, associated to the confined concrete crushing or to the post
tensioning steel rupture.

The principal parameters controlling these states are the position and amount of the
energy dissipation bars and the post tensioning tendons, the initial amount of vertical
load, the section geometry and the initial strain on the post tensioning tendons.
The main drawbacks of these quasi-static tests are the inability to capture the
acceleration spikes in the vertical and horizontal directions due to impact and the inability
to capture the dynamic associated to the system, in fact when rocking is triggered the
system stiffness decreases and therefore the mode of vibration and the lateral load
distribution change.
To better understand the dynamic associated to rocking systems Toranzo (2002) applied
the self centering rocking wall idea to confined masonry constructions. A three story
40% scale confined (by means of reinforced concrete beams and columns) masonry
wall and slab subsystem was tested on a shake table. The tests involved also the use of
steel hysteretic energy dissipation bars between the wall toes and the foundation
element. The test results showed vertical and horizontal acceleration spikes due to wall
impact during rocking, the latter being larger in the upper levels. These spikes lead to
peak absolute accelerations and peak inter-story shear forces higher than expected, in
some cases more than doubled, although the effects on the base shear demand was not
so pronounced and therefore in agreement with the analysis estimation.
The tests showed that the amount of the horizontal acceleration spikes are reduced with
the application of the energy dissipation devices while the vertical ones are not. The
vertical acceleration spikes could temporary reduce the shear friction capacity of the wall
base and lead to a horizontal slip of the wall; this will be analyzed in the next paragraph.
Shake table tests on a rocking wall dragging additional mass have been recently carried
out by Marriott et al. (2008) which studied post tensioned concrete walls with
combinations of hysteretic steel dampers and viscous dampers. The shake table tests
setup was made by a rocking wall connected to the shake table by means of a foundation





ROCKING WALLS IN PRECAST CONCRETE STRUCTURES
175
beam and dragging a crane-suspended concrete mass acting as a pendulum. The
authors studied the experimental response of the post-tensioned walls subjected to
earthquake excitation.
The preliminary results published show the beneficial effects of additional dissipation
devices to the post tensioned wall in damping the response after the main peaks although
the maximum displacement associated to some ground motions could be larger in the
case of additional dissipation devices than without.

11.3 Rocking wall base sliding
In the case of rocking walls the base shear capacity is an important issue, because if it
relies on the shear friction at the wall to foundation joint, this could be not sufficient,
especially when the joint is closed. In fact when the gap opens the elongation of the post
tensioning steel will increase the vertical load and therefore the shear friction capacity.
Yielding in the post tensioning steel will reduce the prestress and therefore the shear
friction capacity. The instant when the gap closes could be critical under a base shear
demand-capacity point of view: the demand could be significant due to impact horizontal
acceleration spikes while the capacity is at minimum due to the low level of vertical load
(the post tensioning force is at minimum, the vertical force in the energy dissipation bars
is acting upward and the vertical acceleration spikes due to impact are acting upward).
A base shear demand greater than the capacity leads to a slip of the base joint with
detrimental effects on the post tensioning tendons and on the energy dissipation bars
which may kink and prematurely failing in the following cycles. Other than that, no re-
centering capacity is associated to horizontal slip which should therefore be avoided.
Knowing the shear friction capacity and demand allows, if necessary, to design shear
resistance passive methods like steel dowels, mechanical keys or socketed connections,
although from a constructability point of view it is better to erect the wall directly on top of
the foundation avoiding additional work related to shear key install. To evaluate if shear
friction capacity is sufficiently large to avoid sliding, Restrepo & Rahman (2007) proposed
a formula based on the effective height to wall depth ratio of the wall. The formula
contains a parameter to take into account the sensitivity of rocking systems to feed high-
frequency energy caused by impact of the wall toes although no value of the parameter
is indicated. To estimate the maximum base-shear demand Kurama et al. (2002) adopted
a formula based on the sum of the first mode component base shear and a higher mode
component which is a function of the first and second mode effective height and mass





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
176
and of the peak ground acceleration. Although the results of time history analyses
showed that the formula adopted provides a good upper bound base shear estimate for
the case study, the equation does not capture the horizontal acceleration spikes, and so
inertia forces, associated to gap closing as rocking occurs: based on that formula,
considering a rocking wall with one single rigid floor (i.e. a single degree of freedom
system), only the first mode base shear component exists with no account of the base
shear demand increase due to the horizontal acceleration peaks as gap closes.
These horizontal acceleration peaks, as recorded in the experimental tests, are more
likely associated to the change in lateral system stiffness when gap closes.
To justify this statement a post tensioned hybrid wall, whose hysteretic energy dissipation
devices are yielding in compression when gap is closing, is considered. The stiffness (k
b
)
of the base moment - rotation (Figure 11.2) when the gap is approaching closure is given
by the contribution of the post tensioning tendons (k
PT
), whose behavior is considered
linear elastic, and the hysteretic energy dissipation bars (k
d
) which, if the gap opening is
big enough, are yielding in compression to allow gap closure.

M
kb

Figure 11.2 Base Moment Rotation relationship

The post tensioning tendons force associated to base rotation is:
0
_
2
w
NA
PT p PT PT
unb PT
L
c
F F E A
l

| |

|
\
= + (11.1)
Where:
F
p0
is the initial prestress
L
w
is the wall depth
l
unb_PT
is the tendons unbonded length





ROCKING WALLS IN PRECAST CONCRETE STRUCTURES
177
c
NA
is the neutral axis
A
PT
is the tendons area
E
PT
is the tendons steel elastic modulus

The tendons moment contribution is:
2 3
w
PT PT
L c
M F
| |
=
|
\
(11.2)
And the tendons stiffness is therefore
2 1
2 1 _
2 3 2
w w PT PT
PT
unb PT
L c L E A M M
k c
l
| || |
= =
| |

\ \
(11.3)
The dissipation bar force associated to base rotation when the dissipation bars are
yielding in compression when the gap is closing is:
( ) ( )
_ _
/ 2
/ 2
y w y
w
d y d d
unb d unb d
L c
L c
F F k E A
l l

= (
(

(11.4)
Where:
F
y
is the yield force
l
unb_d
is the dissipation bars unbonded length
c
y
is the neutral axis as dissipation bars are yielding in compression
A
d
is the dissipation bars area
E
d
is the dissipation bars steel elastic modulus
k is the post yield dissipation bars stiffness

The dissipation bars moment contribution is:
2 3
w
d d
L c
M F
| |
=
|
\
(11.5)
And the dissipation bars stiffness is therefore
2 1
2 1 _
2 3 2
w w d d
d
unb d
L c L E A M M
k k c
l
| || |
= =
| |

\ \
(11.6)
The base moment rotation stiffness is obtained considering that the tendons and the
energy dissipation bars act in parallel:





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
178
_ _
2 3 2
w w PT PT d d
b PT d
unb PT unb d
L c L E A E A
k k k c k
l l
| |
| || |
= + =
|
| |
|
\ \
\
(11.7)
The displacement stiffness associated to the contribution of a floor at a level H
i
is now
considered. This stiffness is obtained considering two systems in series: the first one
(stiffness k
b_
) is the contribution of the base moment rotation relationship considering the
wall acting as a rigid body while the second one (stiffness k
el
) is the wall flexural stiffness
(including shear stiffness will not change the findings) considering the wall base as fixed.
Before gap closes, the first contribution is obtained from the following relations
2
, ,
b
b i b
i i
k
M k F H k F
H H

= = = and therefore:
_ 2
b
b
i
k
k
H

= (11.8)
The second contribution is simply:
3
3
c
el
i
E I
k
H
=
(11.9)
Therefore the lateral displacement stiffness is
_1 _
1 1 1
system el b
k k k

= + ;
_
_1
_
el b
system
el b
k k
k
k k

=
+
(11.10), (11.11)
Once the gap is closed the lateral displacement stiffness is:
_ 2 system el
k k = (11.12)
The stiffness increase once the gap is closed is
( )
2
2 3 2 2 2
_ 2 _
5
_1 _
3 2
3
9
3
3
c b
system el b i i c b
c b
system el b i c i b
i i
E I k
k k k H H E I k
E I k
k k k H E I H k
H H

| |

\
= = =
+ +
+
(11.13)
This stiffness increase when gap closes is associated to an horizontal impact for the
system and therefore explains the horizontal acceleration spikes when gap is closing but
does not explain why these spikes are bigger in the upper floors, as shown in the
experimental tests (Chapter 12). This behavior is explained by the horizontal velocity of
the floor before impact. The floor velocity is associated to the momentum (defined as





ROCKING WALLS IN PRECAST CONCRETE STRUCTURES
179
velocity times mass) which affects the system impulsive response. When gap is closing
(with a velocity ), the upper floors are subjected to a tangential velocity higher than the
one associated to lower floors. This sensibly affects the impulsive response as it is clear
in the experimental tests where the spikes at the upper floors are bigger.

11.4 Equations of motion
The free vibrations of a rocking wall acting as a rigid block have been studied by Housner
(1963). The rigid block is considered oscillating about the centers of rotation O and O
(Figure 11.3)

W
R
O O'
IO
Lw
Hw

Figure 11.3 - Rocking wall free body diagram

When the block is rotated from the vertical by an angle the self weight will exert a
restoring moment. Assuming that there is no horizontal sliding, the equation of motion can
be written as (Housner 1963):
2
2
sin( ) 0
O
d
I WR
dt
+ = (11.14)
Where:
I
O
is the polar moment of inertia about point O
R is the distance between the center of rotation and the wall center of gravity
is the angle between R and the vertical axis
W is the wall self weight






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
180
To simplify the equation to an ordinary differential equation, Housner proposed to
substitute the sine with its argument. A more rigorous approximation related to a small
rocking amplitude is proposed here considering:
sin( ) sin cos sin cos sin cos = = (11.15)
Which leads to:
cos( ) sin( )
O
I WR WR =

(11.16)
The solution of the previous equation is obtained considering the wall released from an
initial rotation
0
and with zero initial velocity:
0
cos( )
tan( ) (tan( ) ) cosh
O
WR
t
I


| |
=
|
|
\
(11.17)
Where cosh(x) is the hyperbolic cosine and it is defined as (e
x
+e
-x
)/2
A full cycle of rocking consists of a wall rotation around O from
0
until reaching the
vertical position, a following rotation around O until reaching
0
(neglecting impact energy
losses) and then back again to the vertical position and
0
rotating around O. The time T
to complete a cycle is the period of free vibration and it is four times the time required to
go from
0
to zero:
1
0
4 tan
cosh
tan cos( )
O
I
T
WR

(11.18)
Once the wall geometry is defined, the period of rocking depends therefore on the initial
gap opening.





ROCKING WALLS IN PRECAST CONCRETE STRUCTURES
181
Based on the work of Housner (1963), Toranzo (2002) exploited the dynamic equation
change when the wall is dragging an additional mass (Figure 11.4).

W
R
O O'
IO
Meff
Lw
heff
Meff heff
Hw

Figure 11.4 Rocking wall dragging additional mass

The results are analogous to the previous equations once I
O
is replaced with
2
_ O comb O eff eff
I I M h = + .

In this research it is proposed to extend the results to systems with additional post
tensioning tendons and energy dissipation devices, Figure 11.5. If the latter are
considered acting in the elastic region, it is still possible to solve the dynamic equation in
close form.

W
O O'
IO
Meff
Lw
Hw
heff
Meff heff
FPT
Fd
Fd

Figure 11.5 Post tensioned hybrid rocking wall dragging additional mass





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
182
The general equations of motion is:
( ) ( )
0
cos h A A B t = (11.19)
And the general rocking period is
1
0
4
cosh
A
T
B A

(11.20)
With:
0
2
_ _
sin( ) 0.5
cos( )
4
p w
w PT PT d d
unb PT unb d
WR F L
A
L A E A E
WR
l l

+
=
| |
+ +
|
|
\
;
2 2
_ _
_
cos( )
4 4
PT PT w d d w
unb PT unb d
O comb
A E L A E L
WR
l l
B
I
+ +
=

(11.21), (11.22)
The results are summarized in the following Table 11.1, in respect of the A and B
parameters.





ROCKING WALLS IN PRECAST CONCRETE STRUCTURES
183

Table 11.1. Rocking and hybrid wall equation of motion parameters
A B

W
R
O O'
IO
Lw
Hw



tan( )

cos( )
O
WR
I

W
R
O O'
IO
Meff
Lw
heff
Meff heff
Hw



tan( )

_
cos( )
O comb
WR
I

W
O O'
IO
Lw
heff
FPT
Hw
Meff
Meff heff



0
2
_
0.5
sin( )
cos( )
4
p w
w PT PT
unb PT
F L
WR
L A E
WR l

+
+



2
_ _ _
cos( )
4
PT PT w
O comb unb PT O comb
WR A E L
I l I

W
O O'
IO
Lw
Hw
heff
FPT
Fd
Fd
Meff
Meff heff



0
2
_ _
0.5
sin( )
cos( )
4
p w
w PT PT d d
unb PT unb d
F L
WR
L A E A E
WR l l

+
| |
+ +
|
|
\


2
_ _ _
2
_ _
cos( )
4
4
PT PT w
O comb unb PT O comb
d d w
unb d O comb
WR A E L
I l I
A E L
l I

+ +
+







DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
184
In the general case of a system subjected to a seismic excitation (
g
u ) the equation of
motion is:
_
cos( ) sgn( )sin( )
2
w
O comb eff eff g
W H
I WR WR M h u
g

| |
= +
|
\

(11.23)

11.5 Design recommendations
In the literature several studies have shown a poly-linear (bi- tri- and tetra-linear)
representation of the base shear lateral displacement relationship (Kurama et al. 1998;
Holden et al. 2003; Restrepo & Rahman 2007). These representations validity has been
proved by the comparison with quasi static cyclic experimental tests consisting of one
lateral load applied at the first mode effective height or a lateral load distribution
corresponding to the first mode distribution. The tests done by those authors allowed to
clearly identify the analytical equations corresponding to the system behavior states. A
set of similar equations have been proposed by Perez et al. (2007) to take into account
the contribution of the inertia forces of each floor to the total base shear and roof drift
once these forces are known without giving indication on how to determine the demand.
An attempt to indicate how to design these systems have been proposed by Kurama et
al. (2002) who suggested to estimate the roof and story drift demands for the design-level
and survival-level ground motions using the equal displacement assumption, which states
that a non linear structure and a linear-elastic structure with the same initial fundamental
period have similar maximum drifts. This assumption is usually applicable for structure
whose fundamental period is greater than 0.5 s and located on stiff soil. The analytical
results showed that the approximation underestimates the effective maximum
displacement and story drift with increasing difference for survival level ground motions,
for wall designed in high seismicity sites and for soft soil sites. To overcome this
drawback Farrow & Kurama (2003) proposed the use of demand index relationships
which, by means of regression analyses, provide the mean values of demand indices like
the peak displacement ductility demand. These relationships have been calibrated for
different hysteretic rules like the bilinear elastic, associated to rocking walls with
additional post-tensioning, and a combined bilinear-elastic/elasto-plastic, suitable for
hybrid walls. This approach is similar to the use of equivalent damping expressions
calibrated for flag shape hysteresis models adopted in the Direct Displacement Based





ROCKING WALLS IN PRECAST CONCRETE STRUCTURES
185
Design (DDBD) procedure. These models well describe the hybrid wall behavior and a
set of the equivalent viscous damping parameters have been calibrated in Chapter 10.
To design rocking walls according to DDBD procedure is possible to consider a linear
deflected shape being the principal source of lateral displacement the wall base gap
opening, especially for low rise buildings less sensitive to higher modes. An alternative
way to evaluate the system equivalent viscous damping for these kind of system has
been proposed by Priestley et al. (2007) by considering the proportion of the total
overturning moment absorbed by the rocking response (M
1
) and by the tensile force
component (M
2
):
1 2
1 2
rocking hysteretic
system
M M
M M

+
=
+
(11.24)
Where:

rocking
is the equivalent viscous damping associated to rocking, usually 0.05.

hysteretic
is the hysteretic damping of the energy dissipation bars as a fraction of their
equivalent viscous damping evaluated with a Jacobsen approach.

11.6 Non linear time history analyses
To model the dynamic behavior of rocking and hybrid wall structures, different
approaches have been proposed. Kurama (2000) compared the use of a finite element
model using Abaqus with a fiber element model using Drain-2dx. The Abaqus finite
element model was made of nonlinear rectangular plane stress elements to model the
wall panels and gap/contact elements to model the gap opening behavior at the base
joint. The Drain-2dx model, simpler from a computational point of view, adopts fiber
beam-column element for the wall, which allowed to develop a reasonably accurate
model using only uniaxial stress-strain models for concrete and post-tensioning steel. To
model gap opening the tensile strength and stiffness of the concrete at the base of the
wall are set to zero. The model accounts for axial-flexural interaction, hysteretic behavior
of mild dissipation steel bars, confined and unconfined concrete, but does not capture the
buckling and low cycle fatigue fracture of the mild steel re-bars nor the additional
degradation in flexural stiffness and resistance of the wall under repeated displacements
at the same amplitude.
The comparison between the two models showed that the fiber wall model is capable of
predicting both the wall global and local behavior, respectively the base shear roof drift





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
186
and the gap opening, therefore has been adopted by those authors for dynamic nonlinear
analyses.
Another way to model rocking structures has been proposed by Toranzo (2002) using
Ruaumoko (Carr 2006). The author used frame elements to model columns and beams.
The steel energy dissipation devices were modeled using the Dodd-Restrepo steel
hysteresis rule and the impact of the base of the wall against the foundation has been
captured by spring elements with a Hertzian Contact Spring hysteresis rule whose
stiffness was defined as the one expected for a concrete strut.
The numerical simulations gave good matching of the experimental tests from the
maximum displacement point of view once the first mode of vibration damping was set to
zero and the damping of other modes set to 3%. In fact when rocking is triggered the first
mode of vibration does not offer significant damping.
A possible way to improve the model could be adding, in series with the contact
elements, non linear springs to capture the confined and unconfined concrete stiffness
and resistance.
Here it is proposed a simpler way to model the rocking mechanism by modeling the base
moment-gap opening relationship with a rotational spring. If properly done this could lead
to a simplified model of the rocking/hybrid wall. The procedure proposed is to get
analytically the main points of the moment-rotation relationship and therefore the
parameters for the flag-shape rotational spring hysteretic model to adopt in the nonlinear
analyses. If the flag shape model is not available among the hysteresis rules, it can be
replaced by two rotational springs in parallel, one with a bilinear elastic hysteretic model
and the other one with an elasto-plastic hysteretic model (Figure 7.4).

Rotation
M
o
m
e
n
t
Rotation
M
o
m
e
n
t
Rotation
M
o
m
e
n
t

Figure 11.6 Flag shape equivalent representation.





ROCKING WALLS IN PRECAST CONCRETE STRUCTURES
187
To determine the base moment-rotation main points it is appropriate to take into account
the confined concrete region behavior and the bond slip between the energy dissipation
bars and the concrete.
The bond-slip model adopted is the one proposed by Zhao & Sritharan (2007), where the
monotonic bar stress () versus loaded end slip (s) is described using a straight line for
the elastic region and a curvilinear portion afterword:
( )
u y y
f f f = + (11.25)
Where:
1/
1
e
e e
R
R R
s
s
s
b s

=
(
| | | |
+ (
| |

\ \ (


( )
/
y y
s s s s =
s
y
slip value when re-bar reaches f
y
s
u
slip value when re-bar reaches f
u

b is the ratio between post yield stiffness and initial elastic stiffness
R
e
is taken as 1.01

The slip at yield of the re-bar is:
( )
1
2.54 2 1 0.34
8437
y
b
y
ck
f
d
s
f

| |
= + + |
|
\
(11.26)
Where the slip and bar dimensions are expressed in mm and the stress in MPa; 0.4 = ,
30 40
u y
s s = and 0.3 0.5 b = .

The confined concrete model proposed by Mander et al. (1983) has been adopted to
determine the confined concrete region parameters, maximum compressive strain (
cu
)
and stress (f
cc
), due to confining steel and concrete arching effect.
The equations ruling the model are the following:





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
188
1 2 cc ck
f f = (11.27)
Where:
1
1.25 1.8 1 7.94 1.6 1
l l
ck ck
F F
f f

| |
= +
|
|
\
;
2
2
1 1.4 0.6 0.8
'
l l l
c l l
F f f
f F F

(
| |
( = +
|
(
\


F
l
is the maximum confining lateral stress.
f
l
is the minimum confining lateral stress.

To take into account the presence of arching encroaching into the concrete core, the
confining lateral stress is reduced according to the effective area of the concrete core :
'
eff
l l
core
A
f f
A
= (11.28)
The effective confined area (A
eff
) is calculated considering that the arching of the concrete
has a parabolic shape with an initial tangent slope of 45 degrees.
The ultimate available compressive strain is
0.004 0.9
300
yt
cu s
f
p = + (11.29)
Where:
Volume of transverse confining reinforcement
Volume of confined concrete core
s
p =
f
yt
yield strength of transverse reinforcement

in the following Figure 11.7 the moment rotation relationship, with the consideration made
before, is shown for the hybrid wall configuration (wall dimensions, post tensioning
tendons amount and initial prestress, energy dissipation bars and confined concrete
parameters) adopted in the experimental tests described in the next chapter.
The left side picture contains the wall response with no additional dissipation devices
while the right side one underline the effect of adding two energy dissipation bars to the
wall. The dash lines show the behavior of the analytical moment rotation relationship
while the diamond shape markers are the experimental results (outlined in Chapter 11).
The analytical relationship slightly overestimates the experimental results and does not
catch the behavior at low rotation values leading to a possible underestimation of the wall
demand; the overall behavior seems reasonable in the case of higher seismicity zones





ROCKING WALLS IN PRECAST CONCRETE STRUCTURES
189
where the base joint opening reaches higher values. Further investigation is required to
compare the non linear time history results of a wall modeled with the simple relationship
proposed to the other finite element model approaches mentioned before. In particular
the model can catch the horizontal acceleration peaks when gap closes but it cannot take
into account the vertical acceleration spikes.

5 0 5
x 10
3
1500
1000
500
0
500
1000
1500
Rotation (rad)
M
o
m
e
n
t

(
k
N
m
)
MomentRotation

5 0 5
x 10
3
1500
1000
500
0
500
1000
1500
Rotation (rad)
M
o
m
e
n
t

(
k
N
m
)
MomentRotation

Figure 11.7 Analytical and experimental moment rotation results.

191
12. SHAKE TABLE TESTS INVOLVING ROCKING WALLS


Rocking and hybrid walls have been recently adopted as lateral force resisting system of
an extensive experimental campaign on precast diaphragms recently concluded at the
University of California, at San Diego and involving shake table tests.
The use of rocking walls compared to classic reinforced concrete walls concentrated the
damage in the floors and allowed the project team to investigate the magnitude of floor
diaphragm forces, load paths and connection behavior in precast buildings. The test
specimen was a three story half scale precast building, designed to be diaphragm
sensitive in flexure (floor aspect ratio of 3.5), with half scale standard precast elements
and typical connection half scale details. The building underwent to extensive shake table
testing to 16 significant input ground motions while 640 sensors recorded dynamically the
behavior of the structure and the details, providing a good opportunity to test a structure
under realistic boundary conditions and allowing the validation of nonlinear finite element
and structural analysis models developed for the project.
In this chapter only the rocking wall results of four significant tests are presented to show
how these systems behave under a real seismic scenario and how they interact with a
real building. Efforts have been made to explain their dynamic behavior in particular
regarding the vertical and horizontal acceleration spikes associated to the wall impacts
once rocking is triggered.
The shake table tests have been carried out with the George B. Brown Jr. Network for
Earthquake Engineering Simulations (NEES) shake table at University of California at
San Diego (UCSD) Englekirk Structural Engineering Center. The shake table has a platen
of 7.6 m wide by 12.2 m long with two servo controlled dynamic actuators with a
combined capacity of 6.8 MN, sufficient to shake the 3.72 MN test structure in the
transverse direction, i.e. exciting the floor diaphragms in their flexible direction.
The simplified building resembled a parking garage and contained three unique floor
systems (17.07x4.88 m): a composite double tee diaphragm at the first floor level, non-
composite hollowcore at the second and pretopped double tee diaphragm at the third.
Floor-to-floor height was 1.98 m; the structure height above the foundation level was 7 m.
The test structure exceeded the shake table footprint and it was therefore built on a





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
192
foundation structure connected to the shake table and extended 4.72 m off each side of
the shake table platen as shown in Figure 12.1.


Figure 12.1 UCSD shake table test setup.





SHAKE TABLE TEST INVOLVING ROCKING WALLS
193
12.1 Hybrid wall general considerations and details
As mentioned before, the lateral force resisting system was made by two rocking walls
whose cross section was 244x20.3 cm and 7.0 m tall. The design of the hybrid walls has
been carried out following the recommendations outlined in Restrepo & Rahman (2007)
where the authors provided analytical equations to describe the bilinear response of the
base shear lateral displacement relationship to use as design check in the dynamic
tests. The authors suggested a bilinear representation to be sufficiently accurate for use
in design and consistent with a two level performance based design: an immediate
occupancy objective, attained at the dissipation bars yield point, and a life safety
performance objective, attained at yielding in the critical post tensioning tendon.
The post-tensioning force, which ensures restoring force and self centering of the rocking
system, has been controlled with the following inequality to ensure tendons remaining
elastic at the life safety performance objective:
( )
0
_
PT PT PT NA
p pp ls
unb PT
E A d c
F F
l

(12.1)
Where:
F
p0
is the initial post-tensioning force
F
pp
is the post tensioning force associated to tendons limit of proportionality
l
unb_PT
is the tendons unbonded length
c
NA
is the neutral axis
A
PT
is the tendons area
E
PT
is the tendons steel elastic modulus
d
PT
is the distance between the furthest tendon to the extreme compression wall toe
fiber.

ls
is the gap opening corresponding to yielding in critical tendon

The post-tensioning has been accommodated with two vertical ducts, each containing five
1/2 (12.7 mm) diameter tendons. Each of the five tendons set was anchored with wedge
anchor plates at the bottom of the foundation beam and at the top of a hollow core
plunger jack (890 kN capacity) placed at the top of the wall. This jack has been used to
seat the tendons (all at once with a force of 565 kN), to record the post-tensioning force
during the tests (allowing the moment base evaluation) and to tailor the post tensioning
force for tests with different seismic demand. All the five strands in each duct were seated





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
194
at once and not one by one as it is usually done in industry practice, this created different
level of initial prestress in each tendon causing a premature failing of the tendons in the
high seismicity tests. Due to the tendons loading setup, once one tendon has failed, the
force exercised by the hollow core plunger jack is transmitted to the other tendons
increasing their stress and therefore the chances of failure in subsequent wall base
rotations. A total of 642 kN was applied to each wall in the larger amplitude tests with an
average strand stress of 0.35-f
pu
, while a prestressing force of 472 kN was applied for
tests in the low seismicity sites due to the lower lateral force demand.
The neutral axis depth, which will define the extension of the confined concrete core, is
found assuming that at life safety performance objective the wall toe concrete cover has
spalled and the vertical force is resisted by the confined concrete region
0.9
'
u pp d
NA
e cc
P F F
c
b f
+ +
(12.2)
Where:
P
u
is the gravity loads on the wall
F
d
is the energy dissipation bars force
b
e
is the confined concrete toe thickness
f
cc
is the confined concrete maximum stress

The confined concrete parameters have been calculated with the model proposed by
Mander et al. (1983) shown in the previous chapter. In each wall toe through thickness
side, a anchor plate has been placed to increase confinement of the concrete, and to
provide a suitable surface to counteract against the shear keys in case of horizontal
sliding of the wall. To avoid the steel plate from carrying compressive forces at impact
instead of the confined concrete (due to the plate higher stiffness) and transmitting the
load to the concrete above the plate with subsequent premature spalling, a Styrofoam
strip has been paced underneath the plate to avoid direct contact with the foundation
grout.
To limit the maximum displacement and damp out the cyclic response during rocking, five
headed #7 (22 mm diameter) re-bars have been placed in corrugated steel ducts
(extending 1.82 m for development length purposes) to act as energy dissipation devices
(Figure 12.2): during rocking the energy dissipation bars elongate as the base gap opens
and therefore dissipate energy in plastic hysteretic response. All the bars have been
grouted in the foundation beam before placing the wall panel.





SHAKE TABLE TEST INVOLVING ROCKING WALLS
195



Figure 12.2 Energy dissipation #7 re-bars South wall

For the first test (low seismicity) the walls have been tested in rocking with no additional
energy dissipation bar; in the following tests (moderate and high seismicity) two energy
dissipation bars for each wall have been grouted symmetrically with respect to the wall
central line.
Placing five dissipation bars in each wall without grouting all of them allowed to replace
the bars prematurely damaged during the test and in a real case scenario to replace them
by grouting the adjacent ones. It is a cost effective procedure to core drill a hole in the
grout between the wall and the foundation to cut the energy dissipation bar in order to
inhibit any further participation in the next tests or seismic events. This operation has
been successfully done during the test campaign to remove an energy dissipation bar
partially grouted (Figure 12.3)





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
196



Figure 12.3 Wall core drilling to disconnect a partially grouted energy dissipation bar.

The total vertical force (gravity load plus post-tensioning force after losses) has been
designed to be sufficient to compress the energy dissipation bars to zero strain upon
unloading:
0 u p d d su
P F F A f + = (12.3)
To limit the maximum dissipation bars strain and set the hysteretic energy equal in all the
energy dissipation bars with no regard to the different positions in the wall cross section,
an unbonded length region is created on the bars. This could be done by a milled portion
or by adding duct tape, as it has been done in the tests. The length of the unbonded
portion (l
unb_d
) is the one associated to a tension strain at the life safety performance
objective (here evaluated as a gap opening of 2%) less than 2/3 of the ultimate
dissipation bar strain.

Another aspect to take into account is the neutral axis length. Restrepo & Rahman (2007)
suggested to limit the neutral axis depth at the life safety performance objective to ensure
both hysteretic response and geometrical stability. They pointed out that the concrete
residual strains at the wall toes for walls displaying a c
NA
/L
w
ratio greater than 0.15
resulted in the loss of initial stiffness upon rocking.
Perez et al. (2003) stated that the neutral axis depth should not interfere with the wall
region with post tensioning ducts (or, referred to this test, un-grouted energy dissipation
bar ducts) because in that case the confined concrete region could separate from the rest





SHAKE TABLE TEST INVOLVING ROCKING WALLS
197
of the wall in this discontinuity region with a consequent buckling of the confined concrete
region. This failure mode has to be avoided because it can jeopardize the wall gravity
load capacity.
Taking into account these considerations, the walls have been designed to limit the c
NA
/L
w

ratio to 0.15 at the life safety performance objective. Particular care has been taken of the
position of shim stacks between the foundation and the wall for erection purposes,
avoiding to place them in the region inside the minimum neutral axis depth.
To design the confinement reinforcement Restrepo & Rahman (2007) suggested that the
ultimate compressive strain of the concrete should be greater than the one corresponding
to a constant strain distribution of length c (neutral axis depth at life safety performance
objective) above the foundation
NA ls
cu ls
NA
c
c

= = (12.4)
This assumption has been adopted in the design of the confined concrete region. Perez
et al. (2007) suggested that the constant strain region should be taken as
{ } min 2 ; 2 "
e NA
b c , where b
e
is the concrete core thickness and c
NA
is the minimum
neutral axis depth without the concrete cover.

Other key aspects taken into account in the wall detailing are:
1. Ensure geometric compatibility between the walls and the floors during rocking.
When rocking is triggered the wall rotates at its toe causing the uplift of the wall to
floor connection, therefore if this connection is monolithic the uplift could tear the
connection apart from the wall jeopardizing the shear transfer mechanism. To
avoid this inconvenience a slotted connection has been successfully adopted
(Figure 12.4) to allow uplift without interfering in the lateral load transmission. The
vertical uplift is a result of the walls flexural response and it is present in
traditionally reinforced concrete cantilever walls as well.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
198



Figure 12.4 Wall to floor connection to allow uplift.

2. Create a sufficiently rough surface between the wall and the foundation to
increase the shear friction capacity (Figure 12.5). This has been done by making
both the contact surfaces of the wall and the foundation beam 3-5 mm rough by
means of a rotor hammer. To increase the toughness of the wall base grout
required by the impact loading during rocking, polypropylene fibers at
approximately 0.02% by weight were added (Figure 12.5).



Figure 12.5 Wall base roughened surfaces and base grout polypropylene fibers.





SHAKE TABLE TEST INVOLVING ROCKING WALLS
199
3. Place shear keys at the wall toes to avoid sliding in the case shear friction
capacity of the wall foundation interface is not enough. At each side of the wall a
shear key has been placed leaving a 2 mm space from the wall edge to allow a
gap opening of about 0.04 rad without contact with the shear key. A steel slug
has been welded to the anchor plate embedded in the wall side to avoid torsion
rotation of the wall base (Figure 12.6).



Figure 12.6 Shear key and steel slug detailed to avoid wall base torsion rotation.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
200
12.2 Hybrid wall design
The walls have been designed according to the considerations shown in the previous
paragraph and following a displacement based design procedure to limit the wall base
joint opening to 3% during the maximum considered earthquake (MCE) for the high
seismic hazard Berkeley (CA) site. With the parameters (energy dissipation bars size,
number of post-tensioning tendons and amount of initial prestress) obtained from this
preliminary design, non linear time history analyses have been carried out with a finite
element model of the whole structure whose results have been taken to refine the design
and to further detail the reinforcement layout. In particular strut and tie models
(Figure 12.7) have been developed to check the steel reinforcement needed in the most
demanding lateral force distribution, obtained from the finite element analyses, for each
floor to floor wall panel, whose results are shown in Figure 12.8.


Figure 12.7 Strut and tie model for the wall lateral force distribution





SHAKE TABLE TEST INVOLVING ROCKING WALLS
201

3
rd
Panel

N
top
= 721 kN
V
top
= 394 kN
M
top
= 16 kNm

2
nd
Panel

N
top
= 1136 kN
V
top
= 428 kN
M
top
= 16 kNm

1
st
Panel

N
top
= 721 kN
V
top
= 394 kN
M
top
= 16 kNm

Figure 12.8 Strut and tie model results.

The design procedure led to the following reinforcement layout (Figure 12.9 Figure 12.10)
and parameters (Table 12.1).






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
202

Figure 12.9 Hybrid wall reinforcement layout: Front view.





SHAKE TABLE TEST INVOLVING ROCKING WALLS
203

Figure 12.10 Hybrid wall reinforcement layout: Cross section AA.

Table 12.1. Reinforcement parameters based on the effective materials adopted.
Unconfined region

Longitudinal reinforcement ratio 0.38 %
Transverse reinforcement ratio 0.34 %
Confined region: cross section dimensions 12x8 (30.5x20.3 cm);
vertical extension 2-6 (76.2 cm)
Longitudinal reinforcement ratio 3.22 %
Transverse reinforcement ratio 1.13 % (Through depth direction)
1.12 % (Through thickness direction)
Maximum confining lateral stress F
l
5.77 MPa
Minimum confining lateral stress f
l
4.69 MPa
Effective area ratio A
eff
/A
core
80 %
Volume of transverse reinforcement
Volume of confined concrete core

2.29 %
Ultimate strain
cu
3.77 %
Confined stress f
cc
84.5 MPa

In the following tables (Table 12.2 Table 12.3) the mechanical characteristics of the wall
concrete, the base grout, the energy dissipation bar (ED) grout and the energy dissipation
steel bars are shown.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
204
Table 12.2. Energy dissipation re-bar properties.
Diameter Yield strength Ultimate strength Peak strain
22.2-mm 490-MPa
a
673-MPa
a
12.3%
b
a
Average of two samples
b
Lowest value obtained from two samples

Table 12.3. Wall concrete and grout properties.
Age Strength Age Strength Age Strength
(days) (MPa) (days) (MPa) (days) (MPa)
Wall concrete 28 54
a
98 51
b
126 55
b

Base joint grout 35 48
b
82 49
b
120 48
b

ED grout in the
foundation beam
28 49
a
87 62
a
115 42
a

ED grout in the wall 13 35
b
NA NA 36 38
b

a
Average of two specimen
b
Average of three specimen

12.3 Test sequence
The input ground motions were selected for three sites in the United States of America
with low (Knoxville-TN), moderate (Seattle-WA) and high (Berkeley-CA) seismic hazard,
the latter one resulting in a near-fault ground motion. The adopted test protocol subjected
the structure to increasingly more demanding input ground motions, ensuring to complete
a sufficient number of tests. The test sequence called for a design basis earthquake
(DBE) for the Knoxville site, followed by a DBE for Seattle, a DBE for Berkeley, and a
maximum considered earthquake (MCE) for the Berkeley site. The representative ground
motions parameters and the scaling factors to match the 5% damped target response
spectra of the relative sites are shown in Table 12.4, while the spectra are shown in
Figure 12.11.





SHAKE TABLE TEST INVOLVING ROCKING WALLS
205
Table 12.4. Full scale structure input ground motions.
Site Earthquake Scaled PGA Scale Factor
Knoxville DBE
TEST 1
1979 Imperial Valley
Parachute Test Site
0.30 1.49
Seattle DBE
TEST 2
1979 Imperial Valley
El Centro Array #5
0.59 1.14
Berkeley DBE
TEST 3
1989 Loma Prieta
Los Gatos Presentation Center
0.41 0.72
Berkeley MCE
TEST 4
1989 Loma Prieta
Los Gatos Presentation Center
0.61 1.08

0 1 2 3 4
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
5
Full Scale Structure Acceleration Spectra
Period (s)
A
c
c

(
g
)


Test 1
Test 2
Test 3
Test 4

Figure 12.11 Full scale structure ground motion spectra.

The scaling of the structure to half scale implied special treatment of the input ground
motions for similitude purposes; all the structural dimensions have been scaled by half
and the same happened to the reinforcement bars diameter and all the connections but
the wall to floor connections, which have been made stronger, through capacity design, to
allow the floor diaphragms study. The purpose of the input ground motions scaling is to
have an half scale structure with the same stresses in the reinforcement bars and in the
connections.
Considering the stresses induced by inertia forces on a connection of area A:
/ / F A ma A = = (12.5)





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
206
In the half scale structure
( )
( )
/ 8
/ 4 2
h h h h
h h
h h
m F m a a m
a
A A A A
= = = = (12.6)
Therefore to maintain the same stresses the input ground motion acceleration amplitude
has to be doubled.
The other aspect to take into account is to make sure that the same relative excitation in
the full scale structure corresponds to the structural mode periods of the half scale
structure: this could be done acting on the input ground motion frequency content.
Considering a cantilever column, with square cross section size B, height H and a mass
on the top; the first mode period is:
3 3
4
2 2 4
3
m mH mH
T
k EI EB
= = = (12.7)
The half scale structure first mode period is
( )( )
( )
3
3 3
4 4 4
/ 8 / 2
2 4 4 2
2
/ 2
h h h
h
h h
m H m m H mH T
T
k EB EB
E B
= = = = = (12.8)
Therefore to maintain the same modes relative excitation the ground motion time step
has to be decreased by half. As a result the seismic event duration is reduced by half.
The half scale structure ground motion spectra are shown in Figure 12.12.

0 1 2 3 4
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
5
Half Scale Structure Acceleration Spectra
Period (s)
A
c
c

(
g
)


Test 1
Test 2
Test 3
Test 4

Figure 12.12 Half scale structure ground motion spectra.





SHAKE TABLE TEST INVOLVING ROCKING WALLS
207
12.4 Instrumentation layout
Six hundred and forty sensors were installed on the building to capture its response. Five
types of sensors were mounted on the structure to monitor accelerations, displacements
or deformations, strains, and pressures. On each wall have been placed:
16 accelerometers, mounted in the direction of shaking and in the transverse and
vertical directions;
13 linear variable displacement transducers (LVDT) to measure the wall uplift and
relative displacement at each floor level;
2 pressure transducers installed on the jacks on top of the walls to measure the
force in the post-tensioning tendons during the tests;
15 strain gages on the wall energy dissipation bars (3 for each bar);
18 concrete strain gages (limited in the North wall - 9 at each toe) to capture
compressive strains at wall rocking (Figure 12.13).

The walls instruments layout is shown in Figure 12.14 and Figure 12.15.



Figure 12.13 North wall East toe concrete strain gages
(the central five strain gages have been placed after Test 2)






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
208

North Wall South Wall


Legend:
Accelerometers LVDT Pressure Transducers

Vertical E-W N-S

Horizontal Vertical

Vertical
1. The accelerometers shown outside the wall are placed on the first precast slab
element adjacent the wall.
2. The N-S accelerometers, placed on the wall sides, monitor the out of plane motion.
3. 0L-42, 0L-43, 0L-44, 0L-45, 0L-46, 0L-47 monitor the base joint opening.
Figure 12.14 Walls instrumentation layout.








SHAKE TABLE TEST INVOLVING ROCKING WALLS
209
North Wall South Wall





Figure 12.15 Walls internal side instrumentation pictures.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
210
12.5 Tests results
In this section the results of the hybrid North wall are shown for the first test of each
selected ground motion (even if more than one test has been conducted for each ground
motion). In Test 1 (Knoxville DBE) no damage was observed. The selected wall post-
tensioning force allowed minor rocking of the wall and no energy dissipation was used
(i.e., the bars were not grouted) at this level of shaking. A maximum 0.33% rotation
occurred at the wall base. Figure 12.16 shows the neutral axis variation at rocking which
varies from 40 to 15% of the wall length when the gap opens.


0
5
10
15
20
25
30
35
40
45
50
N
A

(
%

o
f

L
w
)
Neutral axis variation at rocking


Etoe uplift Wtoe uplift
16 16.1 16.2 16.3 16.4 16.5 16.6 16.7 16.8
0
5
10
U
p
l
i
f
t

(
m
m
)
Time (s)


Etoe Wtoe

Figure 12.16 Test 1 N-Wall neutral axis variation at rocking.

In Figure 12.17 the base moment base joint rotation relationship shows the typical
shape of unbonded post tensioned rocking walls with no additional hysteretic energy
dissipation devices.





SHAKE TABLE TEST INVOLVING ROCKING WALLS
211

3 2 1 0 1 2 3
x 10
3
1500
1000
500
0
500
1000
1500
Base Moment Rotation
(rad)
M

(
k
N
m
)

Figure 12.17 Test 1 N-Wall base moment rotation relationship.

The following two figures (Figure 12.18 Figure 12.19) allow to make comments on the
shear friction capacity and demand at the wall foundation joint. In the first one the wall
vertical acceleration shows that the vertical acceleration spikes happen simultaneously
along the wall height causing a reduction of the vertical force when gap closes and so a
reduction of the shear friction capacity. The slotted wall to floor connection avoided the
vertical acceleration peaks to be transmitted to the adjacent floor precast panels.
The second figure shows the horizontal acceleration (in the direction of excitation) of the
wall and the adjacent precast panels. It is possible to see that the magnitude of the
acceleration spikes when gap closes increases at the upper floors due to the increase of
the horizontal velocity before impact as expected. The figure shows also a delay in the
horizontal acceleration spikes in the precast panels compared to the wall; this delay is
due to the time elapsed between the gap closing and the engagement of the wall to floor
slotted connection. These horizontal acceleration spikes when gap is closing increase the
shear demand on the wall base. Therefore when the base joint gap closes the wall base
shear demand increases and the shear friction capacity is reduced compared to the
condition of a closed gap before rocking is triggered.






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
212

0.5
0.4
0.3
0.2
0.1
0
0.1
0.2
0.3
0.4
0.5
W
a
l
l

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Wall Vertical Acceleration


Wall base Wall top
15.5 16 16.5 17 17.5
0
5
10
U
p
l
i
f
t

(
m
m
)
Time (s)


Etoe Wtoe

Figure 12.18 Test 1 N-Wall vertical acceleration at rocking.

1.5
1
0.5
0
0.5
1
1.5
1
s
t

F
l
o
o
r

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
WallFloor Horizontal Acceleration


Wall 1st slab 2nd slab 3rd slab

1.5
1
0.5
0
0.5
1
1.5
2
n
d

F
l
o
o
r

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)

1.5
1
0.5
0
0.5
1
1.5
3
r
d

F
l
o
o
r

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
16 16.1 16.2 16.3 16.4 16.5 16.6 16.7 16.8
0
5
10
U
p
l
i
f
t

(
m
m
)
Time (s)


Etoe Wtoe

Figure 12.19 Test 1 N-Wall and adjacent precast panels horizontal acceleration.





SHAKE TABLE TEST INVOLVING ROCKING WALLS
213
Figure 12.20 compares the shear demand with the shear friction capacity. The latter is
evaluated for two coefficient of friction (=0.25 and 0.50) to outline possible differences in
the choice of the roughness of the foundation to wall interface. The vertical acceleration
spikes dont affect the wall vertical load too much: the wall is carrying mainly horizontal
loads and the tendon prestress is marginally affected by the vertical acceleration due to
the wall impacts. Regarding the rocking wall shear demand, this has been evaluated
neglecting the contribution in shear transferring of the vertical load system (columns). The
shear demand has been computed as the product between the horizontal acceleration
recorded at each precast panel and the relative tributary mass.
Figure 12.20 shows that for this test the shear demand is well below the capacity.


800
600
400
200
0
200
400
600
800
S
h
e
a
r

(
k
N
)
Shear Demand vs Shear Capacity


Capacity =0.25 Capacity =0.50 Demand
10 15 20 25 30
0
5
10
U
p
l
i
f
t

(
m
m
)
Time (s)


Etoe Wtoe

Figure 12.20 Test 1 N-Wall Base Shear Demand versus Capacity.

Figure 12.21 shows the concrete toe strain increase at impact. The highest values (less
than 0.001, therefore less than the yielding concrete strain in compression) happen at
the wall base layer, showing an almost constant value up to a height corresponding to the
neutral axis length.
An important observation is that the main compressive values are associated to gap
opening rather than wall impacts.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
214


1
0.8
0.6
0.4
0.2
0
x 10
3
1
.
5
"

f
r
o
m

b
a
s
e
(
m
m
/
m
m
)

1
0.8
0.6
0.4
0.2
0
x 10
3
5
.
0
"

f
r
o
m

b
a
s
e
(
m
m
/
m
m
)

1
0.8
0.6
0.4
0.2
0
x 10
3
8
.
5
"

f
r
o
m

b
a
s
e
(
m
m
/
m
m
)

1
0.8
0.6
0.4
0.2
0
x 10
3
1
2
"

f
r
o
m

b
a
s
e
(
m
m
/
m
m
)
Concrete compressive strain


Etoe Wtoe
16 16.1 16.2 16.3 16.4 16.5 16.6 16.7 16.8
0
5
10
U
p
l
i
f
t

(
m
m
)
Time (s)


Etoe Wtoe

Figure 12.21 Test 1 N-Wall concrete strains.

In Test 2 (Seattle DBE) the maximum wall base rotation measured was 0.60%. Two #7
energy dissipation bars were grouted in each wall. The results obtained in this test are
qualitatively similar to those of the previous test. The neutral axis depth reached 15% of
L
w
at maximum gap opening (Figure 12.22). The energy dissipation bars reached a strain
of 0.02 (Figure 12.23), beyond the yield point. The stress values associated to the strain
values recorded have been determined following the Dodd-Restrepo steel hysteretic rule
(Carr 2006). The base moment joint rotation relationship (Figure 12.24) shows the flag
shape typical of the addition of hysteretic energy dissipation devices to rocking walls. It is
clear how the number of energy dissipation bars placed did not compromise the self-
centering capacity of the wall.






SHAKE TABLE TEST INVOLVING ROCKING WALLS
215

10
1
10
0
10
1
0
5
10
15
20
25
30
35
40
Neutral axis depth Rotation
Rotation (%)
N
A

(
%

o
f

L
w
)


Etoe uplift
Wtoe uplift

Figure 12.22 Test 2 N-Wall neutral axis variation at rocking.

0.1 0.05 0 0.05 0.1
800
600
400
200
0
200
400
600
800
Energy Dissipators stressstrain relationship
(mm/mm)


(
M
P
a
)


East ED West ED

Figure 12.23 Test 2 N-Wall Energy Dissipation bars stress-strain relationship
Stress values derived from strains using Dodd-Restrepo formulas (as in Carr 2006).







DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
216
6 4 2 0 2 4 6
x 10
3
1500
1000
500
0
500
1000
1500
Base Moment Rotation
(rad)
M

(
k
N
m
)

Figure 12.24 Test 2 N-Wall base moment rotation relationship.

It is important to note how the energy dissipation bars did not affected either the vertical
(Figure 12.25) and horizontal (Figure 12.26) acceleration spikes at wall impacts once
rocking is triggered, therefore the same considerations of base shear demand and shear
friction capacity at impact apply.


0.5
0
0.5
1
1.5
W
a
l
l

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Wall Vertical Acceleration


Wall base Wall top
13 13.5 14 14.5
0
5
10
15
U
p
l
i
f
t

(
m
m
)
Time (s)


Etoe Wtoe

Figure 12.25 Test 2 N-Wall vertical acceleration at rocking.






SHAKE TABLE TEST INVOLVING ROCKING WALLS
217


1.5
1
0.5
0
0.5
1
1.5
1
s
t

F
l
o
o
r

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
WallFloor Horizontal Acceleration


Wall 1st slab 2nd slab 3rd slab

1.5
1
0.5
0
0.5
1
1.5
2
n
d

F
l
o
o
r

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)

1.5
1
0.5
0
0.5
1
1.5
3
r
d

F
l
o
o
r

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
13 13.5 14 14.5
0
5
10
15
U
p
l
i
f
t

(
m
m
)
Time (s)


Etoe Wtoe

Figure 12.26 Test 2 N-Wall and adjacent precast panels horizontal acceleration.

Figure 12.27 shows the base shear demand versus shear friction capacity: the capacity
reduction associated to the vertical acceleration spikes after impact is still negligible In
this case it is important to make the surface between wall and foundation rough enough
to transfer the shear.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
218


800
600
400
200
0
200
400
600
800
S
h
e
a
r

(
k
N
)
Shear Demand vs Shear Capacity


Capacity =0.25 Capacity =0.50 Demand
10 15 20 25 30
0
5
10
15
U
p
l
i
f
t

(
m
m
)
Time (s)


Etoe Wtoe

Figure 12.27 Test 2 N-Wall Base Shear Demand versus Capacity.

At the maximum base joint rotation (0.60%) a peak concrete compressive strain of
0.00157 was registered at 38 mm from the base grout while an almost constant 0.00085
concrete strain was registered in the wall toe concrete up to 30.5 cm (12) (Figure 12.28).





SHAKE TABLE TEST INVOLVING ROCKING WALLS
219


2
1.5
1
0.5
0
x 10
3
1
.
5
"

f
r
o
m

b
a
s
e
(
i
n
/
i
n
)

2
1.5
1
0.5
0
x 10
3
5
.
0
"

f
r
o
m

b
a
s
e
(
i
n
/
i
n
)

2
1.5
1
0.5
0
x 10
3
8
.
5
"

f
r
o
m

b
a
s
e
(
i
n
/
i
n
)

2
1.5
1
0.5
0
x 10
3
1
2
"

f
r
o
m

b
a
s
e
(
i
n
/
i
n
)
Concrete compressive strain


Etoe Wtoe
13 13.5 14 14.5
0
5
10
15
U
p
l
i
f
t

(
m
m
)
Time (s)

Figure 12.28 Test 2 N-Wall concrete strains.

In Test 3 (Berkeley DBE) the maximum wall base rotation measured was 2.00%. The
same energy dissipation bars from the previous test were used for this test. The data
post-processing showed that an energy dissipation bar in the South wall fractured during
this test and that a 10% drop in post-tensioning force occurred in the same wall. The test
results shown here were unaffected by these changes and went unnoticed, but would be
evidenced in the subsequent test.
The neutral axis depth reached a minimum value of 12% of L
w
(Figure 12.29).
The strain gages of the energy dissipation bars broke during the tests and therefore the
data related to these values, like the wall base moment rotation, are not available.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
220

10
1
10
0
10
1
0
5
10
15
20
25
30
35
40
Neutral axis depth Rotation
Rotation (%)
N
A

(
%

o
f

L
w
)


Etoe uplift
Wtoe uplift

Figure 12.29 Test 3 N-Wall neutral axis variation at rocking.

The vertical (Figure 12.30) and horizontal (Figure 12.31) acceleration spikes at wall
impacts assume in this test significant values, reaching 1.86 g for the vertical acceleration
and 1.5 g for the third floor horizontal acceleration.


1
0.5
0
0.5
1
1.5
2
W
a
l
l

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Wall Vertical Acceleration


Wall base Wall top
13.5 14 14.5 15 15.5 16 16.5
0
20
40
U
p
l
i
f
t

(
m
m
)
Time (s)


Etoe Wtoe

Figure 12.30 Test 3 N-Wall vertical acceleration at rocking.






SHAKE TABLE TEST INVOLVING ROCKING WALLS
221


2
1
0
1
2
1
s
t

F
l
o
o
r

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
WallFloor Horizontal Acceleration


Wall 1st slab 2nd slab 3rd slab

2
1
0
1
2
2
n
d

F
l
o
o
r

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)

2
1
0
1
2
3
r
d

F
l
o
o
r

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
13.5 14 14.5 15 15.5 16 16.5
0
20
40
U
p
l
i
f
t

(
m
m
)
Time (s)


Etoe Wtoe

Figure 12.31 Test 3 N-Wall and adjacent precast panels horizontal acceleration.

Figure 12.32 shows that in this test the wall vertical acceleration spikes at impacts can
sensibly reduce the wall vertical load leading to a shear friction capacity lower than
expected and in some cases lower than the demand. In this study it is outlined how the
use of slotted connection for shear transfer between the wall and the floors is beneficial
because they do not transfer the vertical acceleration spikes on the adjacent floor slabs.
For the same reason it seems appropriate to separate the gravity load system from the
rocking walls, which should be considered and detailed for lateral load transfer system
only, assigning to the post tensioning tendons the self-centering capacity. In this way the
shear friction capacity is maximized.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
222


800
600
400
200
0
200
400
600
800
S
h
e
a
r

(
k
N
)
Shear Demand vs Shear Capacity


Capacity =0.25 Capacity =0.50 Demand
12 14 16 18 20 22 24
0
20
40
U
p
l
i
f
t

(
m
m
)
Time (s)


Etoe Wtoe

Figure 12.32 Test 3 N-Wall Base Shear Demand versus Capacity.

Looking at the concrete strains at impact (Figure 12.34 Figure 12.35), the maximum
compressive strain reached a maximum value of 0.0034 at 5.1 cm from the wall edge.
This is probably due to the position of the concrete strain gages with respect to the
longitudinal reinforcing bars in the confined concrete region (Figure 12.33). The concrete
strain values measured between the re-bars are almost doubled compared to the one
placed close to re-bars.





SHAKE TABLE TEST INVOLVING ROCKING WALLS
223


Figure 12.33 concrete strain gages and longitudinal re-bars position.

The use of headed re-bars as longitudinal reinforcement in the confined region resulted in
a axially stiffer solution than the use of spiral reinforcement. The adopted solution led to
beneficial effects because it helped in carrying the compressive force and limiting the
concrete compressive strains, associated to a very limited damage in the concrete during
the tests.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
224

14 15 16
3
2
1
0
x 10
3
(
m
m
/
m
m
)
(s)

3
2
1
0
x 10
3
(
m
m
/
m
m
)

3
2
1
0
x 10
3
(
m
m
/
m
m
)

3
2
1
0
x 10
3
(
m
m
/
m
m
)
14 15 16
3
2
1
0
x 10
3
(s)
14 15 16
3
2
1
0
x 10
3
(s)

3
2
1
0
x 10
3
14 15 16
3
2
1
0
x 10
3
(s)
14 15 16
3
2
1
0
x 10
3
(s)
13.5 14 14.5 15 15.5 16 16.5
0
20
40
U
p
l
i
f
t

(
m
m
)
Time (s)
West toe uplift

Figure 12.34 Test 3 N-Wall East toe concrete strains at uplift of West toe.





SHAKE TABLE TEST INVOLVING ROCKING WALLS
225

14 15 16
3
2
1
0
x 10
3
(
m
m
/
m
m
)
(s)

3
2
1
0
x 10
3
(
m
m
/
m
m
)

3
2
1
0
x 10
3
(
m
m
/
m
m
)

3
2
1
0
x 10
3
(
m
m
/
m
m
)
14 15 16
3
2
1
0
x 10
3
(s)
14 15 16
3
2
1
0
x 10
3
(s)

3
2
1
0
x 10
3
14 15 16
3
2
1
0
x 10
3
(s)
14 15 16
3
2
1
0
x 10
3
(s)
13.5 14 14.5 15 15.5 16 16.5
0
20
40
U
p
l
i
f
t

(
m
m
)
Time (s)
East toe uplift

Figure 12.35 Test 3 N-Wall West toe concrete strains at uplift of East toe.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
226
Test 4 (Berkeley MCE) was carried out right after Test 3 without discovering the reduced
South wall capacity due to the fractured energy dissipation bar and the post-tensioning
force decrease which affected the test results. Other than that, an energy dissipation bar
fractured also in the North wall once rocking started. This led to a reduced flexural
capacity of the wall which induced extra demand on the post-tensioning tendons causing
their premature failure and the reduction of the post-tensioning force (Figure 12.36). As
said before, tendons failure was likely a result of the method employed to seat and stress
the tendons and occurred at an average strand stress of 0.45-fpu. The ten strands in
each wall were simultaneously seated with hollow core plunger jacks rather than
individually seating each strand as it is done in industry practice, other than that once a
tendon fails the post-tensioning force given by the hollow core plunger jack pressure is
transmitted to the other tendons with their subsequent overload as rocking continues.

0
100
200
300
400
500
600
700
800
P
T

f
o
r
c
e

(
k
N
)
Post tensioning tendons force


Etendon Wtendon
14 15 16 17 18 19
0
100
200
U
p
l
i
f
t

(
m
m
)
Time (s)


Etoe Wtoe

Figure 12.36 Test 4 N-Wall post tensioning force.

The base joint rotation values are not available directly cause the gap opening was
greater than the capacity of the vertical LVDT at the edges of the wall, although a lower
bound estimation is about 4.47%, which has been found by similitude considering the
central wall base vertical LVDT (sensor 0L-43 in Figure 12.13). The neutral axis variation
(Figure 12.37) shows the pure rocking, with no additional post tensioning force and





SHAKE TABLE TEST INVOLVING ROCKING WALLS
227
limited dissipation bar contribution (due to the fracture of one bar), neutral axis depth to
be about 6% of L
w
while the depth before the bar fractured was 14.2% corresponding to a
base joint rotation of 1.84%.
10
1
10
0
10
1
0
5
10
15
20
25
30
35
40
Neutral axis depth Rotation
Rotation (%)
N
A

(
%

o
f

L
w
)


Etoe uplift
Wtoe uplift

Figure 12.37 Test 4 N-Wall neutral axis variation at rocking.

The vertical (Figure 12.38) acceleration spikes at wall impacts reached 3.3 g, directly
related to the gap opening values they were not affected by the bar fracture and the loss
of prestress force.

1
0.5
0
0.5
1
1.5
2
2.5
3
3.5
W
a
l
l

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Wall Vertical Acceleration


Wall base Wall top
14 15 16 17 18 19
0
100
200
U
p
l
i
f
t

(
m
m
)
Time (s)


Etoe Wtoe

Figure 12.38 Test 4 N-Wall vertical acceleration at rocking.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
228
The horizontal (Figure 12.39) acceleration spikes at impact sensibly reduce after the bar
fracture and the loss of post tensioning force, due to the reduction of both the restoring
force and the wall lateral stiffness; these results are in accord with what argued before. It
is possible to note that before bar fracture the horizontal spikes at impact have values
larger than those associated to the maximum gap opening.
The base shear demand could be sensibly higher than expected when the gap closes
and therefore special care is necessary in the design process.


2
1
0
1
2
1
s
t

F
l
o
o
r

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
WallFloor Horizontal Acceleration


Wall 1st slab 2nd slab 3rd slab

2
1
0
1
2
2
n
d

F
l
o
o
r

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)

2
1
0
1
2
3
r
d

F
l
o
o
r

A
c
c
e
l
e
r
a
t
i
o
n

(
g
)
14 14.5 15 15.5 16 16.5 17 17.5 18 18.5 19
0
100
200
U
p
l
i
f
t

(
m
m
)
Time (s)


Etoe Wtoe

Figure 12.39 Test 4 N-Wall and adjacent precast panels horizontal acceleration.






SHAKE TABLE TEST INVOLVING ROCKING WALLS
229
Figure 12.40 shows the reduction of the shear friction capacity once the tendons broke
leading to a shear demand greater than the capacity at wall impacts. The shear not
carried by shear friction has been taken by the shear keys placed at the base of the walls.


800
600
400
200
0
200
400
600
800
S
h
e
a
r

(
k
N
)
Shear Demand vs Shear Capacity


Capacity =0.25 Capacity =0.50 Demand
12 14 16 18 20 22 24
0
50
100
150
U
p
l
i
f
t

(
m
m
)
Time (s)


Etoe Wtoe

Figure 12.40 Test 3 N-Wall Base Shear Demand versus Capacity.

The concrete strains reduced as well due to the post tensioning force loss (Figure 12.41
Figure 12.42).





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
230

14 16 18
3
2
1
0
x 10
3
(
m
m
/
m
m
)
(s)

3
2
1
0
x 10
3
(
m
m
/
m
m
)

3
2
1
0
x 10
3
(
m
m
/
m
m
)

3
2
1
0
x 10
3
(
m
m
/
m
m
)
14 16 18
3
2
1
0
x 10
3
(s)
14 16 18
3
2
1
0
x 10
3
(s)

3
2
1
0
x 10
3
14 16 18
3
2
1
0
x 10
3
(s)
14 16 18
3
2
1
0
x 10
3
(s)
14 15 16 17 18 19
0
100
200
U
p
l
i
f
t

(
m
m
)
Time (s)
West toe uplift

Figure 12.41 Test 4 N-Wall East toe concrete strains at uplift of West toe.





SHAKE TABLE TEST INVOLVING ROCKING WALLS
231

14 16 18
3
2
1
0
x 10
3
(
m
m
/
m
m
)
(s)

3
2
1
0
x 10
3
(
m
m
/
m
m
)

3
2
1
0
x 10
3
(
m
m
/
m
m
)

3
2
1
0
x 10
3
(
m
m
/
m
m
)
14 16 18
3
2
1
0
x 10
3
(s)
14 16 18
3
2
1
0
x 10
3
(s)

3
2
1
0
x 10
3
14 16 18
3
2
1
0
x 10
3
(s)
14 16 18
3
2
1
0
x 10
3
(s)
14 15 16 17 18 19
0
100
200
U
p
l
i
f
t

(
m
m
)
Time (s)
East toe uplift

Figure 12.42 Test 4 N-Wall West toe concrete strains at uplift of East toe.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
232
12.6 Concluding remarks
Despite the wall failure, the structure remained standing. Repairs to the wall were
completed and the testing program was extended for additional tests. The performance of
hybrid walls and building were demonstrated under seismic loading. The walls self
centering capabilities and superior performance benefitted the testing program by
concentrating damage in the floors, the primary focus of the test program, rather than in
the vertical components of the lateral force resisting system.
The results obtained from the experimental program allow to make the following
considerations regarding hybrid wall design. The neutral axis depth was successfully
limited to 15% L
w
by the design procedure. Once the bar fractured and the post
tensioning force was sensibly reduced the neutral axis depth was about 6% of L
w
.
The main contributions of the hysteretic energy dissipation bars were to provide flexural
resistance in the walls without affecting the self centering capacity and to increase the
rocking motion decay after each impact.
The concrete strain increase at impact is negligible compared to the concrete strains due
to the post tensioning force increase associated to tendons elongation when gap opens.
The well confined region with the use of headed longitudinal #5 (15.9 mm diameter)
reinforcement bars provided an axially stiff solution (compared to spiral reinforcement)
which helped in limiting and spreading the concrete strains, which were well below the
design values. Other than that in the design process the axial contribution of the
longitudinal re-bars in the confined region was totally neglected.
At wall impacts, the vertical acceleration spikes are not affected by post tensioning force
or hysteretic energy dissipation devices but are rather depending on the gap opening and
the ground motion intensity. The vertical spikes have the same intensity and happen
simultaneously along the wall height. The horizontal acceleration spikes at impact are
associated to the wall stiffness change once the gap is closed and to the horizontal floor
velocity before impact, with higher values for the upper floors. The horizontal acceleration
spike values at impact are comparable to the maximum horizontal acceleration registered
at the maximum gap opening. As a result the base shear demand and capacity have to
be checked not only at the maximum gap opening (where the shear friction capacity is
maximum due to the increase of vertical force associated to post tensioning tendons
elongation) but at impacts following gap closure; in this latter case the shear friction
capacity is reduced by the vertical acceleration spikes and the demand is increased by
the horizontal acceleration spikes.

233
13. CONCLUSION AND FUTURE DEVELOPMENTS


The primary objective of this thesis has been checking the suitability of the Direct
Displacement Based Design (DDBD) procedure when applied to the seismic design of
precast concrete structures. The research started from the comparison between the
Force Based Design (FBD) and DDBD procedures. At the state of the art this comparison
is not straightforward due to a series of drawbacks in the FBD procedure, the main one is
the failure to predict the inelastic displacements. This is a consequence of the way the
inelastic displacements are related to the structure reduction factor and to the arbitrary
choice of the structural effective stiffness. Once these drawbacks have been overcome,
the two procedures lead to equal results with the difference that FBD computes the
inelastic displacements at the end of the procedure and therefore it is less effective, both
from a logical and computational point of view (because it needs iterations until
convergence), than the DDBD.
Relationships have been derived to relate the hysteretic parameters used to calibrate the
Equivalent Viscous Damping (EVD) equation (used in the DDBD procedure to estimate
the system damping) to the momentcurvature parameters model adopted to describe
the column flexural behavior; this allowed to show that DDBD needs a more rigorous
calibration of the EVD equation especially when applied to structures with moment-
curvature relationships different form the one adopted to determine the EVD parameters
available in the literature. A set of experimental tests has been carried out to analyze the
behavior of some foundation to column connections typical of the precast industry and to
calibrate the parameters of the EVD equation associated to each connection type. A new
calibration algorithm and EVD equation are proposed and applied to different hysteretic
rules; the influence on the ground motion types (i.e. near and far field) and on the
displacement response spectrum shape is evaluated, suggesting a more rigorous EVD
equation definition.
A new equation to relate the yield curvature to the column cross section effective depth
and the axial load ratio is proposed; this formulation overcomes the drawback of the
equation available in the literature especially when applied to some foundation to column
connections typical of the precast industry.





DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
234
The DDBD procedure has been applied to precast structures considering the effects of
the beam to column connections. This allowed to derive specific equations that could be
used in the assessment of existing structures other than in their design. The beam to
column connection contribution leads to a substitute structure, adopted in the DDBD
procedure, with an effective height corresponding to the point of counter flexure and with
an effective mass the whole system mass. The analytical procedure developed has been
applied to the case studies with good results.
Finally the influence of Soil Structure Interaction (SSI) has been taken into account
introducing the foundation flexibility and damping limited to the foundation rocking motion.
This has been done extending results available in the literature which consider a single
degree of freedom substitute structure obtained by static condensation without taking into
account the foundation inertia. The analyses showed that neglecting the foundation
inertia could lead in some conditions to underestimate the system displacements.
Possible developments of this first part of the research are the extension of the analysis
of the DDBD procedure to other precast systems (like multi-story buildings). Regarding
the EVD equation it could be suitable to determine the equation parameters for other
precast foundation to column connections extending the experimental campaign or/and
analyzing the data available in the literature and to investigate the equation modifications
associated to the displacement response spectrum shape.
In the DDBD validation and in the equivalent viscous damping calibration procedure it has
been noted how important is the choice of a suitable set of ground motions whose
characteristics are compatible with the earthquake scenario under interest. A ground
motions scaling procedure is proposed to control and limit the coefficient of variation
(defined as the standard deviation divided by the mean value) of the acceleration
response spectrum of the records set chosen, in order to limit the results variability of non
linear time history analyses and the results dependence on the ground motion bin
chosen. The procedure has been applied to two independent bins of twelve ground
motions each and showed good results in the non linear time history analyses of a non
linear single degree of freedom system with the new scaled records compared to
conventional scaled records. The results obtained in this study could be extended to multi
degree of freedom systems and to single degree of freedom systems with different
hysteretic behavior. The suitability of the procedure in scaling near field records should be
checked as well.







CONCLUSION AND FUTURE DEVELOPMENTS
235


After the application of DDBD procedure to precast structures with classical lateral force
resisting systems (i.e. fixed end columns with or without the contribution of the top column
to beam connection), the use of rocking walls as an alternative resisting system to use in
precast structures has been exploited. This system has self centering properties (given by
post tensioning unbonded tendons) and accommodates the seismic lateral displacement
demand with a base rotation which leads to only one concentrated opening at the
foundation to wall joint compared to the crack spreading and damage typical of the plastic
region of classical reinforced concrete walls. Rocking walls have been the lateral force
resisting system of an extensive experimental campaign on precast diaphragms, involving
shake table tests, recently concluded at the University of California, at San Diego. The
design and the experimental results of these walls have been presented. The
experimental results allowed to give insights on the rocking wall dynamic behavior like the
overestimation of the concrete compressive strain in the design process, and its
negligible increase associated to wall impacts, and the vertical and horizontal
acceleration spikes which affect both the base shear capacity and demand once rocking
is triggered. A moment rotation relationship to use in the non linear analyses and in the
design has been proposed although its suitability has not been exploited yet. The
equations of rocking motions applied to the system under exam have been revisited
especially to determine the rocking period which is believed could play a significant role in
the design procedure although further investigation is required.
The refinement of the design procedure, including the extension of DDBD procedure to
these structures, and recommendations on the detailing are considered as a future
extension of the research.


237
BIBLIOGRAPHY


ASCE 7-05 - Minimum Design Loads for Buildings and Other Structures

Aschheim M. (2002), Seismic Design Based on the Yield Displacement, Earthquake
Spectra, Vol. 18, No. 4, pp. 581-600

Aschheim M., Black E.F. (2000), Yield Point Spectra for Seismic Design and
Rehabilitation, Earthquake Spectra, Vol. 16, No. 2, pp. 317-335

Baker J.W. Cornell C.A. (2005), A Vector-Valued Ground Motion Intensity Measure
Consisting of Spectral Acceleration and Epsilon, Earthquake Engineering and Structural
Dynamics, No. 34, pp. 1193-1217

Baker J.W. Cornell C.A. (2006), Spectral Shape, Epsilon and Record Selection,
Earthquake Engineering and Structural Dynamics, No. 35, pp. 1077-1095.

Blandon C.A. (2004), Equivalent Viscous Damping Equations for Direct Displacement
Based Design, Rose School Master Thesis

Blandon C.A., Priestley M.J.N. (2005), Equivalent Viscous Damping Equations for Direct
Displacement Based Design, Journal of Earthquake Engineering, Vol. 9, Special Issue 2
pp 257-278

Borzi B., Calvi G.M., Elnashaia A.S., Faccioli E., Bommer J.J. (2001), Inelastic Spectra
for Displacement-Based Seismic Design, Soil Dynamics and Earthquake Engineering,
No. 21, pp. 47-61

Bozorgnia Y., Bertero V.V. (2004), Earthquake Engineering: from Engineering
Seismology to Performance Based Engineering, CRC press LLC, New York, USA.






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
238
Browning J.P. (2001), Proportioning of Earthquake-Resistant RC Building Structures,
Journal of Structural Engineering, February 2001, pp. 145-151

Carr A. J. (2006), Ruaumoko, Users manuals, University of Canterbury, Christchurch,
New Zealand

Chopra A.K., Rakesh K.G. (1999), Capacity-Demand-Diagram Methods Based on
Inelastic Design Spectrum, Earthquake Spectra, Vol. 15, No. 4, pp. 637-656

Chopra A.K., Rakesh K.G. (2001), Direct Displacement-Based Design: Use of Inelastic
vs. Elastic Design Spectra, Earthquake Spectra, Vol. 17, No. 1, pp. 47-64

Ciampoli M., Pinto P.E. (1995), Effects of Soil-Structure Interaction on Inelastic Seismic
Response of Bridge Piers, Journal of Structural Engineering Vol. 121, No.5 pp 806-814

Dodd L.L., Restrepo-Posada J.I. (1995), Model for Predicting Cyclic Behaviour of
Reinforcing Steel, ASCE, Vol. 121, No. 3, pp 433-445

Dwairi H.M. (2004), Equivalent Damping in Support of Direct Displacement-Based Design
with Applications to Multi-Span Bridges, PhD Thesis, North Carolina State University,
Raleigh, NC, USA

ETAG 001 (2006), Guideline for European Technical Approval of Metal Anchors for Use
in Concrete

Eurocode 7 (2004), Geotechnical design - Part 1: General rules

Eurocode 8 (2004), Design of Structures for Earthquake Resistance, Part 1: General
rules, seismic actions and rules for buildings

Fajfar P. (2000), A Nonlinear Analysis Method for Performance Based Seismic Design,
Earthquake Spectra, Vol. 16, No. 3, pp. 573-592

Farrow K.T., Kurama Y.C. (2003), SDOF Demand Index Relationship for Performance-
Based Seismic Design, Earthquake Spectra, Vol. 19, No. 4, pp. 799-838





BIBLIOGRAPHY
239

FEMA 274 (1997), Nehrp Commentary on the Guidelines for the Seismic Rehabilitation of
Buildings

FIB Bulletin 25 (2003), Displacement-Based Seismic Design of Reinforced Concrete
Buildings, International Federation for Structural Concrete, Lausanne Switzerland

Freeman S.A. (1998), The Capacity Spectrum Method as a Tool for Seismic Design,
Proceedings of the 11th European Conference on Earthquake Engineering, Paris

Garlock, M.M., Sause, R., and Ricles, J.M. (2007), Behavior and Design of Post-
Tensioned Steel Frame Systems, Journal of Structural Engineering, Vol. 133, No. 3,
pp. 389-399

Gazetas G. (1991) Foundation Vibrations. In Foundation Engineering Handbook, 2nd
Edition, Y.Fang editions, Van Nostrand Reinhold, New York, 553-593

Gazetas G. (1991), Formulas and Charts for Impedance of Surface and Embedded
Foundations, Journal of Geotechnical Engineering Vol. 117, No.9 pp 1363-1381

Gelfi P., Giuriani E., Marini A. (2002). Stud Shear Connection Design for Composite
Concrete Slab and Wood Beams, Journal of structural Engineering, Vol. 128, No. 12, pp
1-7

Ghobarah A. (2001), Performance-Based Design in Earthquake Engineering: State of
Development, Engineering Structures, No. 23, pp. 878-884

Grant D.N., Blandon C.A., Priestley M.J.N. (2004), Modelling Inelastic Response in Direct
Displacement-Based Design, IUSS Press Pavia, Italy

Holden T., Restrepo J., Mander J.B. (2003), Seismic Performance of Precast Reinforced
and Prestressed Concrete Walls, Journal of Structural Engineering, Vol. 129, No. 3,
pp. 286-296






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
240
Housner G.W. (1963), The Behavior of Inverted Pendulum Structures During
Earthquakes, Bulletin of the Seismological Society of America, Vol. 53, No. 2, pp. 403-
417

Jacobsen L.S. (1930), Steady Forced Vibration as Influenced by Damping, Transactions
of ASME 52 pp 169-181

Jacobsen L.S. (1960), Damping in Composite Structures, Proceedings of the Second
World Conference on Earthquake Engineering, pp 1029-1044

Kappos A.J., Manafpour A. (2001), Seismic design of R/C buildings with the aid of
advanced analytical techniques, Engineering Structures, No. 23, pp. 319-332

King D.J., Priestley M.J.N. and Park R. (1986), Computer Programs for Concrete Column
Design, research report 86/12, University of Canterbury, New Zealand.

Kottke A., Rathje E.M. (2008), A Semi-Automated Procedure for Selecting and Scaling
Recorded Earthquake Motions for Dynamic Analyses, Earthquake Spectra, Vol. 24, No.
4, pp. 911-932

Kowalsky M.J. (1994), Displacement-Based Design A Methodology for Seismic Design
Applied to RC Bridge Columns, Master Thesis, University of California San Diego, La
Jolla, CA.

Kurama Y.C., Pessiki S., Sause R., Lu L.-W. (1999), Seismic Behavior and Design of
Unbonded Post-Tensioned Precast Concrete Walls, PCI Journal, May-June 1999, pp. 72-
89

Kurama Y.C. (2000), Seismic Design of Unbonded Post-Tensioned Precast Concrete
Walls with Supplemental Viscous Damping, ACI structural journal, Vol. 97, No. 4, pp. 648-
658

Kurama Y.C. (2002), Hybrid Post-Tensioned Precast Concrete Walls for Use in Seismic
Regions, PCI Journal, September-October, pp. 36-59






BIBLIOGRAPHY
241
Kurama Y.C., Sause R., Pessiki S., Lu L-W (2002), Seismic Response Evaluation of
Unbonded Post-Tensioned Precast Walls, ACI Structural Journal, Vol. 99, No. 5, pp. 641-
651

Mander J.B., Priestley M.J.N., Park R. (1988), Theoretical Stress-Strain Model for
Confined Concrete. ASCE Journal of Structural Engineering, Vol. 114, No. 8

Marriott D., Pampanin S., Bull D., Palermo A. (2008), Dynamic Testing of Precast, Post-
Tensioned Rocking Walls Systems with Alternative Dissipating Solutions, Bulletin of the
New Zealand Society for Earthquake Engineering, Vol. 41, No. 2, pp. 90-103

Montejo L.A., Kowalsky M.J. (2007), CUMBIA Sets of codes for the analysis of
reinforced concrete members, North Carolina State University, Raleigh, NC, USA.

Mukherjee S., Gupta V.K. (2002), Wavelet-Based Generation of Spectrum-Compatible
Time-Histories, Soil Dynamics and Earthquake Engineering, No. 22, pp. 799-804.

Nelder, J.A. and Mead, R. (1965). Downhill Simplex Method on Multi-Dimensional
Minimization, Computer Journal, 7, 308

Panagiotakos T. B., Fardis M. N. (1999), Deformation-Controlled Earthquake-Resistant
Design of RC Buildings, Journal of Earthquake Engineering, Vol. 3, No. 4, pp. 495-518.

Paolucci R., Di Prisco C., Figini R., Petrini L., Vecchiotti M. (2009), Interazione Dinamica
Non Lineare Terreno-Struttura nellAmbito della Progettazione Sismica agli Spostamenti.
Submitted for review to Progettazione Sismica Journal (in Italian)

Paulay T., Priestley M.J.N. (1992), Seismic Design of Reinforced Concrete and Masonry
Buildings, John Wiley and Sons, Inc.

Paulay T. (2002), A Displacement-Focused Seismic Design of Mixed Building Systems,
Earthquake Spectra, Vol. 18, No. 4, pages 689-718






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
242
Paulay T. (2003), Seismic Displacement Capacity Of Ductile Reinforced Concrete
Building Systems, Bullettin of the New Zealand society for Earthquake Engineering,
Vol 36, No. 1

Perez F.J., Pessiki S., Sause R., and Lu L.-W. (2003), Lateral Load Tests of Unbonded
Post-Tensioned Precast Concrete Walls, Special Publication of Large Scale-Scale
Structural Testing, American Concrete Institute, pp. 161-182

Perez, F. J., Sause, R., and Pessiki, S. (2007), Analytical and Experimental Lateral Load
Behavior of Unbonded Post-Tensioned Precast Concrete Walls, Journal of Structural
Engineering, Vol. 133, No. 11, pp. 1531-1540

Pettinga J.D., Priestley M.J.N. (2005), Dynamic Behaviour of Reinforced Concrete
Frames Designed with Direct Displacement-Based Design, IUSS Press, Pavia

Priestley M.J.N. Tao J.R.T. (1993), Seismic Response of Precast Prestressed Concrete
Frames with Partially Debonded Tendons, PCI Journal, Vol. 44, No. 6, pp 42-67

Priestley M.J.N., Seible F., Calvi G.M. (1996), Seismic Design and Retrofit of Bridges,
John Wiley & Sons, Inc., New York, NY, USA

Priestley M.J.N. (1997), Displacement-Based Seismic Assessment of Reinforced
Concrete Buildings, Journal of Earthquake Engineering Vol. 1, No.1 pp 157-192

Priestley M.J.N., Kowalsky M.J. (1998), Aspects of Drift and Ductility Capacity of
Rectangular Cantilever Structural Walls, Bullettin of the New Zealand society for
Earthquake Engineering, Vol. 31, No2

Priestley, M. J. N., Sritharan, S., Conley, J. R. and Pampanin, S. (1999), Preliminary
Results and Conclusions from the PRESSS Five-Story Precast Concrete Test-Building,
PCI Journal, Vol. 44, No. 6, pp. 4267.

Priestley M.J.N. (2002), Direct Displacement-Based Design of Precast/Prestressed
Concrete Buildings, PCI Journal, Nov-Dec 2002






BIBLIOGRAPHY
243
Priestley M.J.N. (2003), Myths and Fallacies in Earthquake Engineering, Revisited, The
Ninth Mallet Milne Lecture, IUSS Press, Pavia

Priestley M.J.N., Grant D.N. (2005), Viscous Damping, in Seismic Design and Analysis,
Journal of Earthquake Engineering, Vol. 9, Special Issue 2 pp 229-255

Priestley M.J.N., Calvi G.M., Kowalsky M.J. (2007), Displacement-Based Seismic Design
of Structures, IUSS Press, Pavia

Restrepo J.I. (2007), SE 223 Advanced Seismic Design, Class notes at University of
California, San Diego - US

Restrepo J.I., Rahman A. (2007), Seismic Performance of Self-Centering Structural Walls
Incorporating Energy Dissipators, Journal of Structural Engineering, Vol. 133, No. 11, pp.
1560-1570

SEOAC (1995). Report On Performance-Based Seismic Engineering, Prepared by the
Vision 2000 Committee of the Structural Engineers Association of California, SEOAC

Shome N., Cornell C.A., Bazzurro P., Carballo J.E. (1998), Earthquakes, Records, and
Nonlinear Responses, Earthquake Spectra, Vol. 14, No. 3, pp. 469-500.

Stewart J.P., Fenves G.L., Seed R.B. (1999), Seismic Soil-Structure Interaction in
Buildings. I: Analytical Methods, Journal of Geotechnical and Geoenvironmental
Engineering, Vol. 125, No 1, pp 26-37

Stone W.C. Cheok G.S. Stanton J.F. (1995), Performance of Hybrid Moment-Resisting
Precast Beam-Column Concrete Connections Subjected to Cyclic Loading. ACI Structural
Journal, Vol. 91, No. 2, pp. 229-249

Sullivan T.J., Priestley M.J.N., Calvi G.M. (2005), Development of an Innovative Seismic
Design Procedure for Frame-Wall Structures, Journal of Earthquake Engineering, Vol. 9,
Special Issue 2, pp 279-307






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
244
Tobolski M.J., Restrepo J.I. (2007), Displacement-Based Design of Cantilever Bridge
Piers, draft paper

Toranzo L.A. (2002), The Use of Rocking Walls in Confined Masonry Structures: a
Performance-Based Approach. PhD thesis, University of Canterbury, New Zealand

Wolf J.P. (1985), Dynamic Soil-Structure Interaction, Prentice-Hall, Inc., Englewood Cliffs,
NJ, USA

Zhao J., Sritharan S. (2007), Modeling of Strain Penetration Effects in Fiber-Based
Analysis of Reinforced Concrete Structures, ACI Structural Journal, Vol. 104, No. 2, pp.
133-141

245
APPENDIX A: SECTION DATA FOR NONLINEAR ANALYSES


The section data for the non linear time histories have been taken from moment-curvature
analyses adopting the characteristic values for materials; the computation uses the
Mander (1988) model for concrete and the Dodd-Restrepo (1995) model for steel.
The ultimate curvature is the curvature associated to the first reaching of: concrete
ultimate strain, steel ultimate strain or moment reduction by 80%.
The equation adopted for the plastic hinge length computation is the one proposed in
Priestley (2007):
0.2 1 0.022
u
p bl y
y
f
L H d f
f
| |
= +
|
|
\

With the notation in Figure 1, the results obtained for square section size ranging from 60
to 110 cm and from 1% to 4% as longitudinal reinforcement ratio are shown in the
following tables.

u
My
ki
r ki
Mu
cr
ki
Mcr
y

Figure 1 Moment curvature data






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
246

Section 60 cm: Plastic hinge length = 0.73m (H=7.9m)

l
M (kNm) (rad/m) M
cr
/M
y
r

1%
Crack 177 0.0005
Yield 650 0.0062 0.27 0.28 0.00032
Ultimate 662 0.1067

2%
Crack 184 0.0004
Yield 964 0.0068 0.19 0.38 0.00008
Ultimate 967 0.0846

3%
Crack 192 0.0004
Yield 1278 0.0072 0.15 0.47 0.00010
Ultimate 1282 0.078

4%
Crack 200 0.0004
Yield 1642 0.0076 0.12 0.46 0.00025
Ultimate 1650 0.0735

Section 70 cm
Plastic hinge length = 0.73m (H=7.9m)

l
M (kNm) (rad/m) M
cr
/M
y
r

1%
Crack 246 0.0003
Yield 876 0.0048 0.28 0.25 0.0009
Ultimate 931 0.091

2%
Crack 255 0.0003
Yield 1282 0.0053 0.20 0.32 0.0008
Ultimate 1325 0.075

3%
Crack 272 0.0003
Yield 1980 0.0059 0.14 0.41 0.00013
Ultimate 1986 0.0667

4%
Crack 285 0.0003
Yield 2477 0.0061 0.12 0.46 0.00036
Ultimate 2496 0.0645





APPENDIX A
247
Section 80 cm
Plastic hinge length = 0.75m (H=7.9m)

l
M (kNm) (rad/m) M
cr
/M
y
r

1%
Crack 332 0.0003
Yield 1307 0.0043 0.25 0.25 0.0011
Ultimate 1413 0.081

2%
Crack 350 0.0003
Yield 2168 0.0049 0.16 0.33 0.00055
Ultimate 2209 0.0622

3%
Crack 369 0.0003
Yield 3068 0.0053 0.12 0.41 0.00024
Ultimate 3086 0.0561

4%
Crack 389 0.0003
Yield 3905 0.0055 0.10 0.47 0.00035
Ultimate 3929 0.0523

Section 90 cm
Plastic hinge length = 0.77m (H=7.9m)

l
M (kNm) (rad/m) M
cr
/M
y
r

1%
Crack 441 0.0002
Yield 1810 0.0037 0.24 0.24 0.0015
Ultimate 2022 0.0742

2%
Crack 469 0.0002
Yield 3225 0.0044 0.15 0.34 0.00048
Ultimate 3278 0.0557

3%
Crack 492 0.0002
Yield 4274 0.0047 0.12 0.40 0.00009
Ultimate 4284 0.0508

4%
Crack 518 0.0002
Yield 5408 0.0049 0.10 0.46 0.00051
Ultimate 5459 0.0463






DISPLACEMENT BASED DESIGN FOR PRECAST CONCRETE STRUCTURES
248
Section 100 cm
Plastic hinge length = 0.77m (H=7.9m)

l
M (kNm) (rad/m) M
cr
/M
y
r

1%
Crack 577 0.0002
Yield 2278 0.0032 0.25 0.23 0.0016
Ultimate 2607 0.0716

2%
Crack 614 0.0002
Yield 4079 0.0038 0.15 0.32 0.00110
Ultimate 4253 0.0525

3%
Crack 649 0.0002
Yield 5786 0.0041 0.11 0.40 0.00009
Ultimate 5798 0.0442

4%
Crack 690 0.0002
Yield 7599 0.0043 0.09 0.47 0.00070
Ultimate 7689 0.0384

Section 110 cm
Plastic hinge length = 0.79m (H=7.9m)

l
M (kNm) (rad/m) M
cr
/M
y
r

1%
Crack 730 0.0002
Yield 2890 0.0028 0.25 0.23 0.0018
Ultimate 3386 0.0661

2%
Crack 780 0.0002
Yield 5399 0.0034 0.14 0.33 0.0012
Ultimate 5675 0.0495

3%
Crack 831 0.0002
Yield 7860 0.0037 0.11 0.41 0.00011
Ultimate 7879 0.0392

4%
Crack 884 0.0002
Yield 10073 0.0039 0.09 0.48 0.00071
Ultimate 10194 0.035

S-ar putea să vă placă și