M.J.N. PRIESTLEY
G.M. CALvI
M.J. KOWALSKY
Displacement-Based.
Seismic Design of Structures
IUSS Press
Istituto Universitario di Studi Superiori di PaviaNessuna parte di questa pubblcseioe pu essere riprnta tasmessa in ais Fema 9 Gon qUaTsas! MEZZO
cletinico, mecca «alto senaa Vautorizeszione seta dei propriate ditt edelleione
No parts of this publication pues be copisd or transmitted 11 any shape or form, and by any type oF electronic.
‘mechanical or diffrent means. Withoul the priot %Hileh permission o tae eopgright holder and the publisher
© Copy right 2007 - 1USS Press
prodote ¢ disiribyitn ds
produced and distrbyed by
Tele (39) 0382.3 7RA1 = fax: (139) 0382.375899
Emit: ini iusspress.it = web: ww jusspress.it
ISBN: 978-88-6198-0000-6very truth passes through three stages (before itis recognized)
In the frst, itis ridiculed
In the second, it is violently opposed
In the third, it is regarded as sof evident”
Arthur Schopenhauer (1788-1860)
“Analysis should be as simple as possible, but no simpler”
Albert Einstein (1879-1955)
‘Strength is essential, but otherwise unimportant”
Hardy Cross! (1885-1959)
" Hardy Cross was the developer of the moment distribution method for structural calculation of
statically indeterminate frames, generally used from the late thirties to the sixties, when it was
superseded by structural analysis computer programs. It seems somehow prophetic that a brilliant
engineer, who based the solution of structural problems on relative stiffness, wrote this aphorism
shat must have sounded enigmatic in the context of elastic analysis and design.CONTENTS
Preface
1 Introduction: The Need for Displacement-Based Seismic Design
1d
12
13
14
Historical Considerations
Force-Based Seismic Design
Problems with Force-Based Seismic Design
1.34 Interdependency of Strength and Stiffness
1.3.2. Period Calculation
1.3.3. Ductility Capacity and Force-Reduction Factors
1.3.4 Ductility of Structural Systems
1.3.5 Relationship between Strength and Ductility Demand
1.3.6 Structural Wall Buildings with Unequal Wail Lengths
1.3.7 Steuctares with Dual (Elastic and Inelastic) Load Paths
1.3.8. Relationship between Elastic and Inclastic
Displacement Demand
1.3.9 Summary
Development of Displacement-Based Design Methods
14.1 Force-Based/Displacement Checked
14.2. Deformation-Calculation Based Design
14.3 Deformation-Specitication Based Design
14.4 — Choice of Design Approach
2 Seismic Input for Displacement-Based Design
21
2.2
23
Introduction: Characteristics of Accelerograms
Response Spectra
22.1 Response Spectra from Accelerograms
2.2.2 Design Elastic Spectra
2.23 Influence of Damping and Ductility on Spectral
Displacement Response
Choice of Accclerograms for Time History Analysis,
3 Direct Displacement-Based Design: Fundamental Considerations
34
3.2
Introduction
Basic Formulation of the Method
v
63
63
63vi
Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
33
34
3.6
37
38
39
3.10
3.2.1 Example 3.1 Basic DDBD
Design Limit States and Performance Levels
3.3.1 Section Limit States
3.3.2 Structure Limit States
3.3.3. Selection of Design Limit State
Single-Degree-of Freedom Structures
3.4.1 Design Displacement for a SDOF seructure
3.4.2 Yield Displacement
3.4.3 Equivalent Viscous Damping
3.4.4 Design Base Shear Equation
3.4.5 Design Example 3.3: Design of a Simple Bridge Pier
3.4.6 Design When the Displacement Capacity Exceeds the
Spectral Demand
3.4.7 Example 3.4: Base Shear for a Flexible Bridge Pier
‘Multi-Degree-of-Freedom Structures
3.5.1 Design Displacement
3.5.2 Displacement Shapes
3.5.3. Effective Mass
3.5.4 Equivalent Viscous Damping
3.5.5 Example 3.5: Effective Damping for a Cantilever Wall Building
3.5.6 Distribution of Design Base Shear Force
3.5.7 Analysis of Structure under Design Forces
3.5.8 Design Example 3.6: Design moments for a
Cantilever Wall Building
3.5.9 Design Example 3.7: Serviceability Design for a
Cantilever Wall Building
P-A Effects
3.6.1 Current Design Approaches
3.6.2. Theoretical Considerations
3.6.3. Design Recommendations for
Direct Displacement-based Design
Combination of Seismic and Gravity Actions
3.7.1 A Discussion of Current Force-Based Design Approaches
3.7.2. Combination of Gravity and Seismic Moments in
Displacement-Based Design
Consideration of Torsional Response in Direct
Displacement-Based Design
3.8.1 Introduction
3.8.2 Torsional Response of Inelastic Eccentric Structures
3.8.3 Design to Include Torsional Effects
Capacity Design for Direct Displacement-Based Design
Some Implications of DDBD
3.10.1 Influence of Seismic Intensity on Design Base Shear Strength
67
67
69
70
72
3
B
75
76
90
o1
92
93
95
96
97
99
100
103
104
105
106,
108
m1
m1
112
14
115
115
119
120
120
122
124
125
127
127Contents vii
3.10.2 Influence of Building Height on Required Frame
Base Shear Strength 129
3.10.3 Bridge with Piers of Different Height 130
3.10.4 Building with Unequal Wall Lengths 132
4 Analysis Tools for Direct Displacement-Based Design 133
4.1. Introduction 133
4.2 Force-Displacement Response of Reinforced Concrete Members 133
4.2.1 Moment-Curvarure Analysis 134
4.2.2 Concrete Properties for Moment-Curvature Analysis 136
4.2.3. Masonry Properties for Moment-Curvature Analyses 139
4.2.4 — Reinforcing Steel Properties for Moment-Curvature Analyses 140
4.25 Strain Limits for Moment-Curvature Analysis 141
4.2.6 Material Design Strengths for
Direct Displacement-Based Design 43
4.2.7 Bilinear Idealization of Concrete Moment-Curvature Curves 144
4.28 — Force-Displacement Response from Moment-Curvature 147
4.2.9 Computer Program for Moment-Curvature and
Force-Displacement 151
4.3 Porce-Displacement Response of Steel Members 151
4.4 Elastic Stiffness of Cracked Concrete Sections, 151
441 Circular Concrete Columns 152
44.2 Rectangular Concrete Columns 155
4.43 Walls 157
4.4.4 Flanged Reinforced Concrete Beams 159
4.45 — Steel Beam and Column Sections 160
4.4.6 Storey Yield Drift of Frames 161
44.7 Summary of Yield Deformations 164
4.5 Analyses Related to Capacity Design Requirements 165
45.1 Design Example 4.1: Design and Overstrength of a Bridge Pier
Based on Moment-Curvature Analysis 167
4.5.2 Default Overstrength Factors 170
45.3 Dynamic Amplification (Higher Mode Effects) 170
4.6 Equilibrium Considerations in Capacity Design 170
4.7 Dependable Strength of Capacity Protected Actions 173
4.7.1 Flexural Strength 173
4.7.2 Beam/Columa Joint Shear Strength 174
4.7.3 Shear Strength of Concrete Members: Modified UCSD model 174
4.7.4 Design Example 4.2: Shear Strength of a Circular
Bridge Column 182
4.7.3. Shear Strength of Reinforced Concrete and Masonry Walls 183
4.7.6 — Response to Seismic Intensity Levels
Exceeding the Design Level 185viii Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures.
48 Shear Flesibility of Concrete Members 185
4.8.1 Computation of Shear Deformations 185
482 Design Example 4.3 Shear Deformation,
and Failure Displacement of a Circular Column 188
4.9 Analysis Tools for Design Response Veritication 192
4.9.1 Introduction 192
49.2 Inelastic Time-History Analysis for Response Veri 192
4.9.3 Non-Linear Static (Pushover) Analysis 218
5 Frame Buildings 221
5.1 Introduction 221
52 ic DDBD Process for Prame Buildings 221
SDOF Representation of MDOF Frame 221
5.2.2 Design Actions for MDOP Structure from,
SDOF Base Shear Force 224
clastic Displacement Mechanism for Frames 225
33 1 Frames 226
53.1 Influence on Design Duetility Demand 226
53.2 stically Responding Frame: 226
5.3.3 Yield Displacement of Irregular Frames 230
53.4 Design Example 5.1: Yield Displacement and
Damping of an Irregular Frame 233
5.3.5 Yield Displacement and Damping when
Beam Depth is Reduced with Height 237
5.3.6 Yield Displacement of Steel Frames 238
34 Controlling Higher Mode Drift Amplification 239
5.5 Structural Analysis Under Lateral Force Vector 242
Analysis Based on Relative Stiffness of Members 242
Analysis Based on Equilibrium Considerations
5.6 Section Flexural Design Considerations 251
6.1 Beam Flexural Design 251
5.6.2 Columa Flexural De: 254
5.7 Direct Displacement-Based Design of Frames for Diagonal Excitation 259
5.8 Capacity Design for Frames 263
General Requirements 263
Beam Flexure 263
5 Beam Shear 265
384 Column Plexure 266
Column Shear 271
5.9 Design Verification 274
5.9.1 Displacement Response 274
5.9.2 Columa Moments
Column ShearsContents ix
5.9.4 Column Axial Forces 27
5.10 Design Example 5.2; Member Design Forces for an
Irregular Two-Way Reinforced Concrete Frame 279
3.11. Precast Prestressed Frames 285
5.11.1 Seismic Behaviour of Prestressed Frames with
Bonded Te 285
3.11.2 Prestressed 287
11.3 Hybrid Precast Beams 290
3.11.4 Design Example 5.3: DDBD of a Hybrid Prestressed
Frame Building including P-A Effects 293
12 Masonry Infilled Frames 301
5.12.1 Structural Options 301
3.12.2. Structural Action of Infill 302
3.12.3. DDBD of Infilled Frames 303
Steel Frames 304
3.13.1. Structural Options 304
5.13.2. Concentric Braced Frames 306
3.13.3 Eccentric Braced Frames 307
5.14 Design Example 5.4: Design Verification of Design Example 5.1/5.2 310
© Scructural Wall Buildings 313
1 Introduction: Some Characteristics of Wall Buildings 313
6.11 Section Shapes 313
6.1.2 Wall Elevations 315
6.1.3. Foundations for Structural Walls 315
6.1.4 Inertia Force Transfer into Walls 317
2 Review of Basic DDBD Process for Cantilever Wall Buildings 317
6.2.1 Design Storey Displacements 317
Wall Yield Displacements: Significance to Desiga 325
63.1 Influence on Design Ductility Limits 325
63.2 Elastically Responding Walls 327
6.3.3 Multiple In-Plane Walls 328
+ Torsional Response of Cantilever Wall Buildings 328
64.1 Elastic Torsional Response 328
6.4.2 Torsionally Unrestrained Systems 331
643 Torsionally Restrained Systems 334
64.4 Predicting Torsional Response 337
64.5 Recommendations for DDBD 339
64.6 Design Example 6.1: Torsionally Eccentric Building 346
6.4.7 Simplification of the Torsional Design Process 352
3 Foundation Flexibility Effects on Cantilever Walls 353
6.3.1 Influence on Damping 353
6.5.2 Foundation Rotational Stiffness
354Priestley, Calvi and Kowalsky.
isplacement-Based Seismic Design of Structures
6.6 Capacity Design for Cantilever Walls
6.6.1 Modified Modal Superposition (MMS) for
Design Forces in Cantilever Walls
& 66.2 Simplified Capacity Design for Canuilever Walls
6.7 Precast Prestressed Walls
68 Coupled Structural Walls
6.8.1 General Characteristics
6.8.2 Wall Yield Displacement
68.3 Coupling Beam Yield Drift
6.8.4 Wall Design Displacement
6.8.5 Equivalent Viscous Damping
6.8.6 Summary of Design Process
68.7 Design Example 6.3: Design of a Coupled—Wall Building
Dual Wall-Frame Buildings
7.1 Introduction
7.2. DDBD Procedure
7.2.1 Preliminary Design Choices
7.2.2. Moment Profiles for Frames and Walls
7.2.3. Moment Profiles when Frames and Walls are
Connected by Link Beams
7.24 — Displacement Profiles
7.2.5 Equivalent Viscous Damping
7.2.6 Design Base Shear Force
7.2.7 Design Results Compared with Time History Analyses
7.3 Capacity Design for Wall-Frames
7.3.1 Reduced Stiffness Model for Higher Mode Effects
7.3.2. Simplified Estimation of Higher Mode Effects for Design
7A — Design Example 7.1: Twelve Storey Wall-Frame Building
7.4.1 Design Data
4.2 Transverse Direction Design
7.4.3 Longitudinal Direction Design
7.44 Comments on the Design
Masonry Buildings
8.1 Introduction: Characteristics of Masonry Buildings
8.1.1 General Considerations
8.1.2 Material Types and Properties
8.2 Typical Damage and Failure Modes
8.2.1 Walls
8.2.2 Coupling of Masonry Walls by Slabs, Beams or
Masonry Spandrels,
8.3. Design Process for Masonry Buildings
357
359
363
370
372
372
376
378
379
381
382
382
387
387
388
388
389
392
304
396
397
397
399
400
401
403
403
404
410
441i
43
413,
413
415
418
418
425
429xi
3 Timb
we Bridg
it
33.1 Masonry Coupled Walls Response
83.2 Design of Unreinforced Masonry Buildings
8.3.3. Design of Reinforced Masonry Buildings
3D Response of Masonry Buildings
84.1 Torsional Response
84.2 Out-of-Plane Response of Walls,
yer Structures
Introduction: Timber Properties
Ductile Timber Structures for Seismic Response
9.2.1 Ductile Moment-Resisting Connections in Frame Construction
9.2.2 Timber Framing with Plywood Shear Panels
9.2.3 Hybrid Prestressed Timber Frames
DDBD Process for Timber Structures
Capacity Design of ‘Timber Structures
es
Introduction: Special Characteristics of Bridges
10.1.1 Pier Section Shapes
10.1.2. The Choice between Single-columa and Multi-column Piers
10.1.3 Bearing-Supported vs. Monolithic Pier/Superstructure
Connection
10.1.4 Soil-Structure Interaction
10,1.5 Influence of Abutment Design
10.1.6 Influence of Movement Joints
10.1.7 Multi-Span Long Bridges
10.18 P-A Effects for Bridges
10.1.9 Design Verification by Inelastic Time-History Analyses
Review of Basic DDBD Equations for Bridges
Design Process for Longitudinal Response
10.3.1 Pier Yield Displacement
10.3.2 Design Displacement for Footing-Supported Piers
10.3.3 Design Example 10.1; Design Displacement for a
Footing Supported Column
10.34 Design Displacement for Pile /Columns
10.3.5 Design Example 10.2: Design Displacement for a Pile/Column
10.3.6 System Damping for Longitudinal Response
10.3.7 Design Example 10.3: Longitudinal Design of a
Four Span Bridge
Design Process for Transverse Response
10.4.1 Displacement Profiles
10.4.2 Dual Seismic Load Paths
10.4.3 System Damping
429
439
446
446
449
455
487
460
457
460
461
462
463
465
465
465
467
467
468
470
470
470
471
471
47
472
472
478
481
483
484
485
489
494
495
498
498xii Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
10.4.4 Design Example 10.4: Damping for the Bridge of Fig. 10.17 500
10.4.5. Degree of Fixity at Column Top 502
10.4.6 Design Procedure 503
" 10.4.7 Relative Importance of Transverse and Longitudinal Response 505
10.4.8 Design Example 10.5: Transverse Design
of a Four-Span Bridge 507
10.5 Capacity Design Issues 512
10.5.1 Capacity Design for Piers 512
10.5.2 Capacity Design for Superstructures and Abutments 513
10.6 Design Example 10.6: Design Verification of Design Example 10.5 516
11 Structures with Isolation and Added Damping 519
11.1 Fundamental Concepts 519
11.1.1 Objectives and Motivations 519
11.1.2 Bearing Systems, Isolation and Dissipation Devices 522
11.1.3 Design Philosophy/Performance Criteria 523
11.1.4 Problems with Force — Based Design of Isolated Structures $24
11.1.5 Capacity Design Concepts Applied to Isolated Structures 526
11.16 Alternative Forms of Artificial Isolation/ Dissipation 527
11.1.7 Analysis and Safety Verification 528
11.2 Bearing Systems, Isolation and Dissipation Devices 529
11.2.1 Basic Types of Devices 529
11.22 “Non-Seismic” Sliding Bearings 530
11.2.3. Isolating Bearing Devices 531
11.2.4 Dissipative systems 544
11.25 Heat Problems 554
11.2.6 Structural Rocking as a Form of Base Isolation 557
11.3 Displacement-Based Design of Isolated Structures 559
11.3.1 Base—Isolated Rigid Structures 559
11.3.2 Base-Isolated Flexible Structures 571
11.3.3 Controlled Response of Compiex Structures 579
11.4 Design Verification of Isolated Structures 596
11.4.1 Design Example 11.7: Design Verification of
Design Example 11.3 596
11.4.2 Design Example 11.8: Design Verification of
Design Example 11.5 597
12 Wharves and Piers 599
12.1 Introduction 599
12.2 Structural Details 601
12.3. The Design Process 602,
12.3.1. Factors Influencing Design 602
123.2. Biaxial Excitation of Marginal Wharves 603Consens
xiii
12.3.3 Sequence of Design Operations
24 Port of Los Angeles Performance Criteria
12.4.1 POLA Earthquake Levels and Performance Criteria
12.4.2 Performance Criteria for Prestressed Concrete Piles
12.4.3 Performance Criteria for Seismic Design of Steel Pipe Piles
Lateral Force-Displacement Response of Prestressed Piles
Prestressed Pile Details
Moment-Curvature Characteristics of Pile/Deck Connection
Moment-Curvature Characteristics of Prestressed
Pile In-Ground Hinge
12.5.4 Inelastic Static Analysis of a Fixed Head Pile
Design Verification
12.6.1 Eccentsicity
12.6.2 Inelastic Time History Analysis
27 Capacity Design and Equilibrium Considerations
12.7.1 General Capacity Design Requirements
12.7.2 Shear Key Forces
Design Example 12.1: Initial Design of a Two-Segment Marginal Wharf
2.9 Aspects of Pier Response
45 _Displacement-Based Seismic Assessment
Introduction: Current Approaches
13.4.1 Standard Force-Based Assessment
13.1.2 Equivalent Elastic Strength Assessment
13.1.3 Incremental Non-linear Time History Analysis
Displacement-Based Assessment of SDOF Structures
13.2.1 Alternative Assessment Procedures
13.2.2 Incorporation of P-A Effects in Displacement-Based
Assessment
13.2.3 Assessment Example 13.1: Simple Bridge Column
under Transverse Response
Displacement-Based Assessment of MDOF Structures
13.3.1 Frame Buildings
13.3.2 Assessment Example 2: Assessment of a
Reinforced Concrete Frame
13.3.3. Structural Wall Buildings
13.3.4 Other Structures
24 Draft Displacement-Based Code for Seismic Design of Buildings
rences
Sembols List
604
608
609
609
oil
612
612
613
618
621
628
628
630
634
634
638
639
645
647
647
649
649
650
653
653
656
659
661
666
672
676
677
61
703xiv Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Abbreviations 113,
indéx 15
Structural Analysis CD 721,PREFACE
formance-based seismic design is a term widely used by, and extremely popular
the seismic research community, but which is currently rather irrelevant in the
ce of design and construction. In its purest form, it involves a large number of
Dilistic considerations, relating to variability of seismic input, of material properties,
Eamensions, of gravity loads, and of financial consequences associated with damage,
2 “apse or loss of usage following seismic attack, amongst other things. As such, it is a
“Soult tool to use in the assessment of existing structures, and almost impossible 0 use,
' expectation of realism, in the design of new structures, where geometry becomes
variable, and an almost limitless number of possible design solutions exists.
srently, probability theory is used, to some extent, in determination of the seismic
which is typically based on uniform-hazard spectra, However, structural engineers
‘nis information and design structures to code specified force levels which have been
_steemined without any real consideration of risk of damage or collapse. Structural
= s>lacements, which can be directly related to damage potential through material strains
rural damage) and drifts (non-structural damage), are checked using coarse and
sole methods at the end of the design process. At best, this provides designs that
=stv damage-control criteria, but with widely variable risk levels. At worst, it produces
segs of unknown safety,
This text attempts to bridge the gap between current strucrural design, and a full (and
ssibly unattainable) probabilistic design approach, by using deterministic approaches.
! on the best available information on analysis and material properties to produce
sures that should achieve, rather than be bounded by, a structural or non-structural
== state under a specified level of seismic input. Structures designed to these criteria
at be termed “uniform-risk” structures. The approach used is very simple —
~svalent in complexity to the most simple design approach permitted in seismic design
the “equivalent lateral force procedure”), but will be unfamiliar to most designers,
- design displacement is the starting point. The design procedure determines the
‘hear force, and the distribution of strength in the structure, to achieve this
placement. The process (displacements lead to strength) is thus the opposite of current
sign, where strength leads to an estimate of displacement, Although this requires a
in thinking on the part of the designer, it rapidly becomes automatic, and we
ssueve, intellectually satisfying,
xvxvi Priestey, Calvi and Kowalsky. Displacement-Based Scismic Design of Structures
This book is primarily directed cowards practising structural designers, and follows
from two earlier books with which the principal author has been involved (“Seismic
Design of Reinforced Concrete and Masonry Structures” (with T. Paulay), John Wiley,
1992, and “Seismic Design and Retrofit of Bridges” (with F. Scible and G.M. Calvi), John
Wiley, 1996). ‘These books primarily address issues of section design and detailing, and 10
a limited extent force-distribution in the class of structures addressed, Although great
emphasis is given in these books to seismic design philosopay in terms of capacity design
considerations, comparatively little attention is directed towards an examination of the
optimum level of strength required of the building or bridge. This text addresses this
aspect specifically, but also considers the way in which we distribute the required system
strength (the base-shear force) through the structure. This takes two forms: methods of
structural analysis, and capacity desiga, It is shown that current analysis methods have a
degree of complexity incompatible with the coarseness of assumptions of member
stiffness, Frequently, equilibrium considerations rather than stiffaess considerations can
lead to a simpler and more realistic distribution of strength. Recent concepts of inelastic
torsional response have been extended and adapted to displacement-based design.
Combination of gravity and seismic effects, and P-A effects are given special
consideration.
Capacity design considerations have been re-examined on the basis of a large number
of recent research studies. Completely new and more realistic information is provided for
a wide range of structures. Section analysis and detailing are considered only where new
information, beyond that presented in the previous ewo texts mentioned above, has
become available.
The information provided in this book will be of value, not just to designers using
displacement-based principles, but also to those using more conventional force-based
design, who wish to understand the seismic response of structures in more detail, and to
apply this understanding to design.
Although the primary focus of this book is, as noted above, the design profession, it is
also expected to be of interest to the research community, as it provides, to our
knowledge, the firs attempt at a complete design approach based on performance
criteria. A large amount of new information not previously published is presented in the
book. We hope it will stimulate discussion and further research in the area. The book
should also be of interest to graduate and upper-level undergraduate students of
earthquake engineering who wish to develop a deeper understanding of how design can
be used to control seismic response.
‘The book starts with a consideration in Chapter 1 as to why ie is necessary to move
from force-based to displacement-based seismic design. ‘This is largely related to the
guesses of initial stiffness necessary in force-based design, and the inadvisability of using
these initial stiffness values to distribute seismic lateral force chrough the structure.
Chapter 2 provides a state-of-the-science of seismic input for displacement-based design,
particularly related to characteristics of elastic and inelastic displacement spectra. ‘The
fundamental concepts behind “direct displacement-based seismic design” — so-called
because no iteration is required in the design process — are developed in Chapter 3.« specially relevant to displacement-based design are discussed in Chapter
sanciples of displacement-based design are then applied to different structural
walls, dual wall/frames, masonry and timber buildings, bridges,
seismic isolation and added damping, wharves) in the following chapters,
uently adapted to seismic assessment in Chapter 13. Finally, the principles
in Chapter 14 in a code format to provide a possible basis for furure
The text is illustrated by design examples throughout.
gn procedure outlined in this book has been under development since first
ceccesi 5 the early 1990's, and is now in a rather complete form, suitable for design
Much of the calibration and analytical justification for the approach has
considerable number of research projects over the past five or so years, and
consequently wish to particularly acknowledge the work of Juan Camillo
eiandro. Amaris, Katrin Beyer, Carlos Blandon, Chiara Casarotti, Hazim
0 Grant, Pio Miranda, Juan Camillo Ortiz, Didier Pettinga, Dario Pietra,
z, and Tim Sullivan, amongst others.
Tas Jesign verification examples described in the book have been prepared with the
= Rui Pinho, Dario Pietra, Laura Quaglini, Luis Montejo and Vinicio Suarez.
T analysis software employed in such design verifications has been kindly
= cof. Athol Carr, Dr. Stelios Antoniou, Dr. Rui Pinho and Mr. Luis Montejo,
me LS: agreed to make these programs available in the Structural Analysis CD.
seple who need special acknowledgement are Prof. Tom Paulay and Dr. Rui
=5o cach read sections of the manuscript in draft form, noted errors and made
for improvements. Their comments have significantly improved the final
remaining errors are the responsibility of the authors alone. Advice from
Bommer, Prof. Ezio Faccioli, and Dr. Paul Somerville on aspects of
<= and of Prof, Guido Magenes on masonry structures is also gratefully
ancial assistance of the Italian Dipartimento della Protegione Cinile, who funded 2
ject on displacement based design coordinated by two of the authors, is
acknowledged.
sy, Christebarch
Pavia1
INTRODUCTION: THE NEED FOR DISPLACEMENT-
BASED SEISMIC DESIGN
1 HISTORICAL CONSIDERATIONS
“arthquakes induce forces and displacements in structures. For elastic systems these
Jirectly related by the system stiffness, but for structures responding inelastically, the
= conship is complex, being dependent on both the current displacement, and the
Displacement Displacement
(a) Structure (b) Profiles at Yield (c) Base Shear/Displacement
Response
Fig.19 Influence of Foundation Flexibility on Displacement Ductility Capacity18 Priestley, Calvi and Kowalsky. Displacemem-Based Seismic Design of Structures
increase, due to strain hardening of longitudinal reinforcement, This minor effect is
ignored, in the interests of simplicity, in the following,
‘The similarity t0 the case of the previous example of the portal frame is obvious. By
analogy to the equations of that section, the displacement ductility of the wall, including
foundation Mexibility effects can be related to the rigid-base case by
— Ae ty Ho!
a,+a, | (eA,/a,
My =1+ (1.19)
where Ay and Ay are the wall displacements at yield due to structural deformation of the
wall, and foundation rotation respectively, and fla, = 1+Ay/Ay.
‘The reduction in displacement duclity capacity implied by F.q(1.19) is more eritical
for squat walls than for slender walls, since the flexural component of the structural yield
displacement, which normally dominates, is proportional to the square of the wall height,
whereas the displacement due co foundation flexibility is directly proportional to wall
height. It is not unusual, with squat walls on spread footings, to find the displacement
ductility capacity reduced by a factor of two or more, as a consequence of foundation
totation effects, Similar effects have been noted for bridge columns on flexible
foundations. ‘To some extent, however, the effects of additional elastic displacements
resulting from this cause may be mitigated by additional clastic damping provided by soil
deformation and radiation damping!“"1, For simplicity, shear deformation of the wall has
not been considered in this example.
In the past it has been common for designers to ignore the increase in fundamental
period resulting from the foundation flexibility discussed above. It may be felt that this to
some extent compensates for the reduction in displacement ductility capacity, since the
structure is designed for higher forces than those corresponding to its “true” fundamental
period. However, the consequence may be that story drifts exceed codified limits without
the designer being aware of the face.
(@) Structures wich Unequal Column Heights: Marginal wharves (wharves running
parallel to the shore line) typically have a transverse section characterized by a simple
reinforced or prestressed concrete deck supported by concrete or stecl shell pile/columns
whose free height between deck and dyke increases with distance froin the shore. An
‘example is shown in Fig.1.10(a).
Conventional force-based design would sum the clastic stiffnesses of the different
piles to establish 2 global scructural stiffness, calculate the corresponding fundamental
period, and hence determine the clastic lateral design force, in accordance with the
sequence of operations defined in Fig.1.3. A force-reduction factor, reflecting the
assumed ductility capacity would then be applied to determine the seismic design lateral
force, which would then be distributed between the piles in proportion to their stiffness.
Implicic in this approach is the assumption of equal displacement ductility demand for all
pile/columns.Chapter 1, Introduction: The Need for Displacement-Based Seismic Design »
The illogical nature of this assumption is apparent when the individual pile/columa
force-displacement demands, shown in Fig.1.10(b), are investigated. Design is likely to be
such that only one, or at most ewo pile designs will be used, varying the amount of,
prestressing or reinforcing steel, but keeping the pile diameter constant. In this case the
pile/colamns will all have the same yield curvatures, and yield displacements will be
proportional to the square of the effective height from the deck to the point of effective
fixity for displacements, at a depth of about five pile diameters below the dyke surface.
‘This effective height is shown for piles F and C in Fig. 1.10(a) as Hy or He.
Concrete Deck
t
ry
He
F rE Ip ic B ix
(2) Transverse Section through Wharf
: F
a
E
D
c
B
A
Or Tr Displacemént
(©) Force-Displacement Response of Individual Piles
Fig.1.10 Transverse Seismic Response of a Marginal Wharf
The structure lateral force displacement response can be obtained by summing the
individual pile/column force-displacement curves, shown in Fig.1.10(b). Force-based20 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
design, allocating strength in proportion to the elastic stiffnesses would imply design
strengths for the different pile/columns equal to the forces intersected by the line drawn
in Fig.1.10(b) at Ayr, the gield displacement of pile/column F. Since the yield
displacements of the fonger piles are much greater, the full strength of these piles will
thus be under-utilized in the design. It is also clearly a gross error to assume that all piles
will have the same ductility demand in the design-leve! earthquake. Fig. 1.10(b) includes,
the full force-displacement curves, up to ultimate displacement, for pile/columns F and
E. The ultimate displacements for the longer pile/columns are beyond the edge of the
graph. Clearly at the ultimate ductility capacity of the shortest pile column, F, the ductility
demands on the longer columns are greatly reduced. Pile/columns A, and B will still be in
the elastic range wien the ultimate displacement of pile F is reached. The concept of a
force-reduction factor based on equal ductility demand for ail pile/columns is thus totally
inapplicable for this structure. Wharf seismic design is discussed in depth in Chapter 12.
Similar conclusions (as well as a means for rationally incorporating the above within
the framework of force-based design) have been reached by Paulayl%! referring to
response of a rigid building on flexible piles of different lengths. The procedure suggested
by Paulay requires thae the concept of a specified structural force-reduction factor, which
currently is a basic tenet of codified force-based design, be abandoned, and replaced by
rational analysis.
A second example, that of a bridge crossing a valley, and hence having piers of
different heights, is shown in Fig.1.11. Under longitudinal seismic response, the
deflections at the top of the piers wili be equal. Assuming @ pinned connection berween
the pier tops and the superstructure (or alternatively, fixed connections, and a rigid
superstructure), force-based design will allocate the seismic design force between the
columns in proportion to their elastic stiffnesses. If the columns have the same cross-
section dimensions, as is likely to be the case for architectural reasons, the design shear
forces in the columns, V4, Va,and Ve, will be in inverse proportion to Hi), Hy? and He
respectively, since the stiffness of column 7 is given by
K,=C,El,,(H} (1.20a)
where Ae is the effective cracked-section stiffness of column & typically taken as 0.5 Jpros
for all columns. The consequence of this design approach is that the design moment at
the bases of the piers will be
M,,=C,V,H, =C,C,EI,,/H?, (1.206)
that is, in inverse proportion to the square of the column heights (in Fgs.(1.20), G and
G are constants dependent on the degree of fixity at the pier top). Consequently the
shortest piers will be allocated much higher flexural reinforcement contents than the
longer piers. This has three undesirable effects. First, allocating more flexural strength to
the short piers will increase their elastic flexural stiffness, Ee, even further, with respect
to the mote lightly reinforced longer piers, as has been discussed in relation to Fig.1-4Chapter 1. Introduction: The Need for Displacement-Based Seismic Design 2
8
Fig.1.11 Bridge with Unequal Column Heights
A redesign should strictly be carried out with revised pier stiffnesses, which, in
accordance with Eq, (1.20) would result in still higher shear and moment demands on the
shorter piers. Second, allocating a large proportion of the total seismic design force to the
short piers increases their vulnerability to shear failure. Third, the displacement capacity
of the short piers will clearly be less than that of the longer piers. As is shown in Section
1.3.5, the displacement capacity of heavily reinforced columns is reduced as the
longitudinal reinforcement ratio increases, and hence the force-based design approach
will tend to reduce the displacement capacity
As with the marginal wharf discussed in the previous example, the ductility demands
on the piers will clearly be different (inversely proportional to height squared), and the
use of a force-reduction factor which does not reflect the different ductility demands will
clearly result in structures of different safety.
Design of bridges with unequal column heights is considered further in Chapter 10.
1.3.5. Relationship between Strength and Ductility Demand
‘A common assumption in force-based design is that increasing the strength of a
structure (by reducing the force-reduction factor) improves its safety. The argument is
presented by reference to Fig.1.1, of which the force-deformation graph is duplicated
here as Fig:1.12{@). Using the common force-based assumption that stiffness is
independent of strength, for a given section, it is seen that increasing the strength from
SI to S2 reduces the ductility demand, since the final displacement remains essentially
constant (the “equal displacement” approximation is assumed), while the yield
displacement increases. It has already been noted, in relation to Fig.14 that this
assumption is not valid. However, we continue, as it is essential to the argument that
increasing strength reduces damage.
The reduction in ductility demand results in the potential for damage also being
decreased, since structures are perceived to have a definable ductility demand, and the
lower the ratio of ductility demand to ductility capacity, the higher is the safety.
We have already identified three flaws in this reasoning: 1) stiffness is not independent
of strength; 2) the “equal displacement” approximation is not valid; and 3) it is not
possible to define a unique ductility capacity for a structural type22 Priestley, Calvi and Kowalsky. DisplacementsBased Seismic Design of Structures
Strength,
14a
isplacement
Reinforcement Ratio (%)
(@) Strength vs Ductility (b) Influence of Rebar % on Parameters
Fig.1.12 Influence of Strength on Seismic Performance
Itis of interest, however, to examine the argument by numeric example. ‘The simple
bridge pier of Fig. 1.1 is assumed to have the following properties; Height = 8 m (26.2 ft),
diameter = 1.8 m (70.9 in), flexural reinforcement dia. = 40 mm (1.58 in), concrete
strength P, = 39 MPa (5.66 ksi), flexural reinforcement: yield strength 4 = 462 MPa (67
ksi), & = 1.54; transverse reinforcement: 20 mm (0.79 in) diameter at a pitch of 140 mm
6.3 in), fy = 420 MPa (60.9 ksi); cover to main reinforcement = 5G mm (1.97 in), axial
load P= 4960 KN (1115 kips) which is an axial (oad ratio of P/Padg = 0.05.
A reference design with 1.5% flexural reinforcement is chosen, and analyses carried
out, using the techniques described in Chapter 4 to determine the influence of changes to
the flexural strength resulting from varying the flexural reinforcement ratio between the
limits of 0.5% and 4%. Results are presented for different relevant parameters in
Fig.1.12(b) as ratios to the corresponding parameter for the reference design.
As expected, the strength increases, almost linearly with reinforcement ratio, with
ratios between 0.5 times and 2.0 times the reference strength. We can thus use these data
to investigate whether safety has increased as strength has increased. First we note that
the effective stiffness has not remained constant (as assumed in Fig,1.12(a)) but has
increased at very nearly the same rate as the strength, More importantly, we note that the
displacement capacity displays the opposite trend from that expected by the force-based
argument: that is, the displacement capacity decreases as the strength increases. At a
reinforcement ratio of 0.5% it is 31% higher than the reference value, while at 4%
reinforcement ratio the displacement capacity is 21% lower than the reference value.
“Thus, if the “equat displacement” approach was valid, as illustrated in Fig.1.12(a), we have
decreased the safety by increasing the strength, and we would be better off by reducing
the strength.
OF course, the discussion above is incomplete, since we know that the yield
displacements are not proportional to strength, since the stiffness and strength are closelyChapter 1, Introduction: The Need for Displacement-Based Seismic Design 23
related as suggested in Fig. 1.4(b), and demonstrated in Fig.1.12(b). We use this to
determine the influence on displacement ductility capacity, and find that it decreases
slightly faster than the displacement capacity (sce Fig.1.12(b)). However, since the clastic
stiffness increases with strength, the elastic period reduces, and the displacement demand
is thus also reduced. If we assume that the structural periods for all the different strength
levels lie on the constant-velocity slope of the acceleration spectrum (je. the linear
portion of the displacement response spectrum: see Fig1.2(b)), then since the period is
proportional to the inverse of the square root of the stiffness (Eq.1.6), the displacement
demand will also be related to 1/4°5, We can then relate the ratio of displacement
demand to displacement capacity, and compate with the reference value.
This ratio is also plotted in Fig.1.12(b). It will be seen that taking realistic assessment
of stiffness into account, the displacement demand/capacity ratio is insensitive to the
strength, with the ratio only reducing from 1.25 to 0.92 as the strength ratio increases by
400% (corresponding to the full range of reinforcement content). Clearly the reasoning
behind the strength/safety argument is invalid.
4.3.6 Structural Wall Buildings with Unequal Wall Lengths
A similar problem with force-based design to that discussed in the previous section
occurs when buildings are provided with cantilever walls of different lengths providing
seismic resistance in a given direction. Force-based design to requirements of existing
codes will require the assumption that che design lateral forces be allocated to the walls in
proportion to their elastic stiffess, with the underlying assumption that the walls will be
subjected to the same displacement ductility demand. Hence the force-reduction factor is
assumed to be independent of the structural configuration,
It was discussed in relation to Fig.1.4(b), that the yield curvature for a given section is
essentially constant, regardless of strength. It will be shown in Section 4.4.3 that the form
of the equation governing section yield curvature is
9, =C-Esh (1.21)
where his the section depth, and & is the yield strain of the longitudinal reinforcement.
Since the yield displacement can be related to the yield curvature by Eq.(1.13) for
cantilever walls, as well as for columns, it follows that the yield displacements of walls of
different lengths must be in inverse proportion to the wall lengths, regardless of the wall
strengths. Hence displacement ducility demands on the walls must differ, since the
maximum response displacements will be the same for each wall
Figure 1.13 represents 4 building braced by two short walls (A and C) and one long
wall (B) in the direction considered. The form of the force-displacement curves for the
walls are also shown in Fig.1.13. Force-based design mistakenly assumes that the shorter
walls can be made to yield at the same displacement as the longer wall B, and allocates
strength beeween the walls in proportion to 4, since the elastic stiffnesses of the wall
differ only in the value of the wall effective moments of inertia, f, which are propostional24 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Force
cig Le peal tlic lee Ae Det cone?
Fig.1.13 Building with Unequal Length Cantilever Walls
to the cube of wall length. Again stcength is unnecessarily, and unwisely concentrated in
the stiffesc elements, underutilizing the more flexible members. A more rational decision
would be to design the walls for equal flexural reinforcement ratios, which would resule in
strengths proportional to the square of wall length.
As with the previous two examples, the code force-reduction factor for the structure
will not take cognizance of the fact that the different walls must have different
displacement ductility demands in the design carthquake.
1.3.7, Structures with Dual (Elastic and Inelastic) Load Paths.
A more serious deficiency of force-based design is apparent in structures which
possess more than one seismic load path, one of which remains elastic while the others
respond inelastically at the design carthquake level. A common example is the bridge of
Fig.1.14(a), when subjected to transverse seismic excitation, as suggested by the double-
headed arrows. Primary seismic resistance is provided by bending of the piers, which are
designed for inelastic response, However, if the abutments are restrained from lateral
displacement transversely, superstructure bending also develops. Current seismic desiga
philosophy requires the super-structure to respond elasticallyl™l, The consequence is that
a portion of the seismic inertia forces developed in che deck is transmitted to the pier
footings by column bending (path 1 in Fig.1.14(b)), and the remainder is transmitted as
abutment reactions by superstructure bending (path 2). Based on an elastic analysis the
relative elastic stiffnesses of the two load paths are indicated by the two broken lines in
Fig.1.14(b), implying that column flexure (path 1) carries most of the seismic force. A
force-reduction factor is then applied, and design forces determined
‘The inelastic response of the combined resistance of the columns is now shown by
the solid line (path 3, in Fig.1.14(b)), and on the basis of the equal displacement
approximation it is imagined that the maximum displacement is Amas, the value predicted
by the elastic analysis. If the superstructure is designed for the force developed in path 2Chapter 1. Introduction: The Need for Displacement-Based Seismic Design 25
al
Column 77 |
Pianig
{-Stper
“ “pees?
3
DH
N
ml
Force
‘Column,
Inelastic
A, Displacement Amax
(a) Structure (b) Load-path Characteristics
Fig.1.14 Bridge with Dual Load Paths under Transverse Excitation
at the column yield displacement, it will be seriously under-designed, since the forces in
this path, which are required to be within the elastic range, continue to rise with
increasing displacement. Thus the bending moment in the superstructure, and the
abutment reactions at A and E are not reduced by column hinging, and a force-reduction
factor should not be used in their design.
It is also probable that the maximum response displacement will differ significantly
from the initial elastic estimate, since at maximum displacement, the effective damping of
the system will be less than expected, as hysteretic damping is only associated with load
path 3, which carries less than 50% of the seismic force at peak displacement response in
this example. This may cause an increase in displacements. On the other hand, the higher
strength associated with the increased post-yield stiffness of load path 2 may result in
reduced displacement demand. Elastic analysis and the force-reduction factor approach
give no guidance to these considerations.
A slightly different, but related problem occurs with dual wall/ frame buildings (see Fig,
Displacement
(@) Seructure (b) Force-displacement Response
Fig.1.15 Dual Wall/Frame Building26 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
1,15). If the seismic force is distributed between the frame and the wall in proportion to
their clastic stiffness, the load-carrying capacity of the frame will be unnecessarily
discounted. ‘The yield displacement of the frame will inevitably be several times larger
thafi that of the wall, so the proportion of seismic force carried by the frame at maximum
response will be larger than at first yield of che wail (Fig, 1.14(b)). In this example both
systems eventually cespond inelastically, but the frame system remains elastic to larger
displacements,
Note that the interaction between the frame and wall due to resolving the
incompatibilities between their natural vertical displacement profiles wil! also be modified
by inelastic action, and bear little resemblance to the elastic predictions. This is discussed
farther in Chapter 7.
1.3.8 Relationship between Elastic and Inelastic Displacement Demand
Force-based design requires assumptions to be made when determining the maximum
displacement response. The most common assumption is the equal-displacement
approximation, which states that the displacement of the inelastic system is the same as
that of the equivalent system with the same clastic stiffness, and unlimited strength (refer
to Fig,1.1). Thus, with reference to Fig.1.2, the design displacemenc is estimated as
T
axles BE Ark
Arnaxcuctie =A (1.22)
and hence # = R, Equation (1.22) is based on the approximation that peak
displacements may be related to peak accelerations assuming sinusoidal response
equations, which is reasonable for medium period structures.
‘The equal displacement approximation is known to be non-conservative for short-
period structures. As a consequence, some design codes, notably in Central and South
American, and some Asian countries, apply the equal-energy approximation when
determining peak displacements. The equal energy approach equates the energy absorbed
by the inelastic system, on a monotonic displacement to peak response, to the energy
absorbed by the equivalent clastic system with same initial stiffness. Thus the peak
displacement of the inelastic system is
f
1\_ 7?
A =A WAH)
rmax,duct vaxelastic \ QR } 4n
R41)
ne SE] (1.23)
where R is the design force-reduction factor. Since Amaxelasic =RAy, and the actual
displacement ductility demand is = Amasauc/Ay, the ductility demand implied by
Bq.(1.23) isChapter 1. Introduction: The Need for Displacement-Based Seismic Design 21
He (1.24)
Where codes employ inelastic design spectra [eg. X1], design is based on specified
ductility, rather than force-reduction factor, and the design spectral accelerations for
short-period structures are adjusted to correct for displacement amplification.
In the United States, where until recently the dominant building code for seismic
regions has been the UBC [X5], design displacements were estimated as
4,38
A,
mmaraer = Ay
(1.25)
where A, is the yield displacement corresponding to the reduced design forces, found
from structural analysis. Since the structure is designed for a force-reduction factor of R,
this would appear to imply that the displacement ductility is
=k
8
and the displacement of the ductile system is 3/8 of the equivalent elastic system.
However, the apparent reason behind this seemingly unconservative result is that the
actual force-reduction factor was substantially lower than the design force-reduction
factor, as a consequence of the design period being pegged to an unrealistic heighr-
dependent equation of the form of Eq.(1.7). The consequences of this are explained with
reference to Fig. 1.16.
(1.26)
08
z
06
od
02
0
Period (sec)
Fig.1.16 Influence of Under-predicted Period on Actual Force-Reduction Factor
In Fig.1.26, TL and T2 are the fundamental periods corresponding to the code height-
dependent equation, and rational seructural analysis respectively. The elastic response
accelerations corresponding to these periods are a1 and a respectively. If the design28 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
force-reduction factor corresponding to T1 is R, then the real force-reduction factor,
corresponding to T2 is R = Raz/ar. If the equal displacement approximation were valid
at T2, and assuming a constant velocity range far che response spectrum, then Eq,(1.26)
would be correct if 2 = 2.67 T1. Fxamination of Table 1.1 indicates that this is close to
the ratio of periods calculated by rational analysis and by the period dependant code
equation,
Clearly there are compensatory errors involved in this approach, which should be
removed by using more realistic periods, and force-reduction factors that have a close
relationship to the ductility capacity, as is incorporated in other codes, Recently, USA.
practice, incorporated in the IBC codel*t, has changed, with the 38/8 factor of Eq,(1.25)
being replaced by 2 coefficient dependent on structural form and material. The approach
is, however, still illogical, with effective ratios varying between 0.5Rand 1.2R,
A compatison of the different predictions provided by Eqs (1.22), (1.24) and (1.26) is
presented in Fig. 1.17, for a design force-reduction factor of R=4, The range of different
possible answers is disturbingly large.
Force
Elastic 77 Europe,NZ,
Japan, Sth
IBC'06 “America
je(R4N /2
ay Aa Displacement
Fig.1.17 Estimates of Design Displacement from Different Force-Based Codes
forR=4
Force-based seismic design does not normally take account of the different hysteretic
characteristics of different materials and structural systems. Thus the fact that seismic
isolation systems absorb much more hysteretic energy than reinforced concrete
structures, which in turn absorb more than prestressed concrete structures is not directly
considered, though different force-reduction factors may be assigned to different
materials. Figure 1.18 examines the validity of the equal displacement approximation for a
range of different periods, and for three different hysceresis rules: elastic, bilinear elasto-
plastic (representative of isolation systems), Takedal0'l, (representative of reinforced
concrete structures, and flag-shaped with B=0.35 (representative of hybrid unbonded
prestressed structures - see Fig 4.33).
Analyses were first carried out for a range of periods between 0.25sec and 2.5 seconds
using elastic time-history analyses and a suite of seven accelerograms compatible with theChapter 1. Introduction: The Need for Displacement-Based Seismic Design 29
FC8 design spectrum for firm ground [X3], The design yield strength for the three ductile
systems was found by dividing the average maximum clastic response moment by a factor
of R=4, All three ductile systems adopted the same force-displacement envelope, with a
post-yield stiffness of 5% of the initial stiffness, and thus only differed in terms of
unloading and reloading ules. Elastic damping was taken as 5% of critical, related to the
tangent stiffness (sce Section 4.9.2(g) for a discussion on modelling elastic damping).
In Fig.1.18, results are expressed as the ratio ductile peak displacement for ductile
response to displacement of the clastic system of equal initial period. For the equal
displacement approximation to hold, al values should be 1.0. It is seen that significant
differences occur, depending on the period and hysteresis rule. Differences are
particularly marked in the period range T <0,75 seconds, as expected, but are also
significant at other periods
25
Displacement Ratio (\/A.))
Displacement Ratio (A/c)
05 + T T T To 08 TT T 1
Cn ee 0 05 4 ass
Period (seconds) Period (seconds)
(a) Absolute Peak Displacement (b) Average of Positive and Negative Peaks
Fig.1.18 Ratio of Ductile to Elastic Peak Displacement for different Hysteresis
Rules based on EC8 Design Spectrum for Firm Ground
The differences between the hysteretic mules, and also from the clastic results are
particularly apparent when the average of the positive and negative peaks (Fig. 1.18(6)),
rather than the absolute maximum (Fig. 1.17(a)), are considered, as suggested in Section
4.9.2(h), reflecting the larger residual displacements in the Bilinear elasto-plastic results
which affects the absolute peak displacement more than the average of positive and
negative peaks.
‘The results of this brief section indicate that force-based design is not ideaily suited to
estimating the maximum displacements expected of structures in seismic response.
Considering that it is now accepted that peak displacements are critical in determining the
level of damage that can be expected, this is a serious criticism of the method.30 Priestiey, Calvi and Kowalsig, Displacement-Based Seismic Design of Structures
13.9 Summary
In this section, we have identified some of the problems associated with force-based
desfgn, These can be summarized as follows:
© Force-based design relies on estimates of initial stiffness 10 determine the period
and the distribution of design forces between different structural elements. Since
the stiffness is dependent on the strength of the elements, this cannot be known
until the design pracess is complete.
© Allocating scismic force between elements based on initial stiffness (even if
accurately known) is illogical for many structures, because it incorrectly assumes
that the different elements can be forced to yield simultaneously.
© Force-based design is based on the assumption that unique force-reduction
factors (based on ductility capacity) are appropriate for a given structural type
and material. This is demonstrably invalid.
Despite these criticisms it should be emphasized that current force-based seismic
design, when combined with capacity design principles and careful detailing, generally
produces safe and satisfactory designs. However, the degree of protection provided
against damage under a given seismic intensity is very non-uniform from structure to
structure, ‘Thus, the concept of “uniform risk” which is implicit in the formulation of
current seismic design intensity, has nat been continued into the structural design. We
believe that it should be.
1.4. DEVELOPMENT OF DISPLACEMENT-BASED DESIGN METHODS
1.4.1 Force-Based/Displacement Checked
Deficiencies inherent in the force-based system of seismic design, some of which have
been outlined in the preceding sections, have been recognized for some time, as the
importance of deformation, rather than strength, in assessing seismic performance has
come to be better appreciated. Consequently a number of new design methods, or
improvements to existing methods, have been recently developed. Initially the
approaches were designed to fit within, and improve, existing force-based design. These
can be characterized as fore-based/ displacement checked, where enhanced emphasis is placed
on realistic determination of displacement demand for structures designed to force-based
procedures.
Such methods include the adoption of more realistic member stiffnesses for
deformation (if not for required strength) determination, and possibly use of inelastic
time-history analysis, or pushover analysis, to determine peak deformation and drift
demand. In the event that displacements exceed the code specified limits, redesign is
required, as suggested in Fig. 1.3, Many modern codes | e.g. X1, X2, X3, X4]require some
version of this approach. Several recent design approaches have used this approach (e.g:
F1, F2, X8}. In general, no attempt is made to achieve uniform risk of damage, or of
collapse for structures designed to this approach.Chapter 1. Introduction: The Need for Displacement-Based Seismic Design 31
Paulay! has suggested that the deficiencies noted in previous sections can be
eliminated within a force-based design approach. «As explained in detail in section 4.4,
yield displacement Ay can be determined from section and structure geometry without a
prior knowledge of strength, Displacement demand, Ag, at least for fame buildings will
normally be governed by code drift limits and the building geometry. ‘The yield strength V
is assumed, and hence the initial stiffness K'= V/A, is calculated. The elastic period is
calculated from Eq,(1.6), and the elastic displacement demand from Eq, (1.22). This is
compared with the code drift limit, and the strength adjusted incrementally until the
elastic displacement equals the drift limit. Strengch is then distributed between the
different lateral-force resisting elements based on experience, rather than on elastic
stiffness. This has been termed a displacement facused force-based approach.
There are, however, problems associated with this approach. Although the yield
displacements of the lateral-force resisting elements may be known at the start of the
procedure, the equivalent system yield displacement will not be known until the
distribution of strength between elements is decided. The approach relies on assumptions
about the equivalence between clastic and ductile displacements (eg. the equal
displacement approximation), which as discussed in relation to Fig.1.18 may be invalid,
and considerable experience is required of the designer. The procedure is suitable for
those well versed in seismic design, but ill-suited for codification. As will be shown in
subsequent chapters of this text, a design approach based directly on displacements is
simpler, better suited co codification (see Chapter 14), and does not require assumptions
to be made about elastic /inclastic displacement equivalence.
1.4.2, Deformation-Calculation Based Design
‘A more refined version of the force-based/ displacement-checked approach sclates the
detailing of critical sections (in particular details of transverse reinforcement for
reinforced concrete members) to the local deformation demand, and may hence be
termed deformation-caleulation based design. Stength is related to a force-based design
procedure, with specified force-reduction factors. Local deformation demands, typically
in the form of member end rotations or curvatures are determined by state-of-the-art
analytical tools, such as inelastic pushover analyses or inelastic time-history analyses.
‘Transverse reinforcement details are then determined from state-or-the-art relationships
benveen transverse reinforcement details and local deformation demand, such as those
presented in Chapter 4
Initial work on this procedure was related to bridge structures! and followed by
work on reinforced concrete buildings("!, Many additional variants of the approach have
recently been developed [e.g. B1, K1, P7]. In the variant suggested by Panagiatokos and
Fardis!™™ the structure is initially designed for strength to requirements of direct
combination of gravity load plus a serviceability level of seismic force, using elastic
Per, comm. T-Paulay32 Priestley, Calvi and Kowalsky. Displacememt-Based Seismic Design of Structures
analysis methods. The designed structure is then analysed using advanced techniques such
as inclastic time-history analysis or inelastic pushover analysis to determine the required
transverse reinforcement details. It is not clear that this is an efficient design approach
when response to the full design-level earthquake is considered, since inelastic time-
history analyses of frame buildings by Pinto et al | 51 have indicated that member
inelastic rotations are rather insensitive to whether gravity loads arc incorporated in the
analysis, or ignored. An alternative procedure for combination of gravity and seismic
loads is suggested in Section 3.7. The approaches described in this section have the
potential of producing structures with uniform risk of collapse, but not with uniform risk
of damage.
1.4.3 Deformation-Specification Based Design
Recently a number of design approaches have been developed where the aim is to
design structures so that they achieve a specified deformation state under the design-level
earthquake, rather than achieve a displacement that is less than a specified displacement
limit. ‘These approaches appear more philosophically satisfying than those of the
preceding two sections. This is because damage can be directly related to deformation.
Hence designing structures to achieve a specified displacement limit implies designing for
4 specified risk of damage, which is compatible with the concept of uniform risk applied
to determining the design level of seismic excitation. It thus means that different
structures designed to this approach will (jdeally) have the same tisk of damage, rather
than the variable risk associated with current design approaches, a8 discussed in Section
1.3. Using state-of-the-art detailing/deformation relationships, structures with uniform
tisk of collapse, as well as of damage can theoretically be achieved.
Different procedures have been developed to achieve this aim, The most basic
division between them is on the basis of stiffness characterization for design. Some
methods [e.g Al, C2, St], adopt the initial pre-yield elastic stiffness, as in conventional
force-based design. Generally some iteration is required, modifying initial stiffness and
strength, to achieve the desired displacement, as discussed in relation to the approach
suggested by Paulay in Section 1.4.1, These approaches also rely on existing relationships
between elastic and inelastic displacement, such as the equal-displacement, or equal-
energy approximations. It is shown in Section 4,9.2(g) that these approximations have
been based on invalid clastic damping assumptions.
“The second approach utilizes the secant stiffness to maximum displacement, based on
the Substitute Structure characterization! and an equivalent elastic representation of
hysteretic damping at maximum response (e.g. P8, K2, P9|. Generally these methods
require little or no iteration to design a structure to achieve the specified displacement,
and are hence known as Direct Displacement-Basea Design (DDBD) methods. The different
stiffness assumptions of the two approaches are illustrated for a typical maximum
hysteretic force-displacement response in Fig.1.19, where Kj and K, are the initial and
secant stiffness to maximum response respectively. It will be recalled that one of the
principal problems with force-based seismic design is that reliance on initial stiffnessChapter 1. Introduction: The Need for Displacement-Based Seismic Design 33
results in illogical force distribution between different structural elements. It will be
shown in Chapter 3 that this problem disappears when the secant stiffness is used.
Displacement
First cycle
‘Subsequent cycle
Fig.1.19 Initial and Secant Stiffness Characterization of Hysteretic Response
‘The way in which hysteretic energy dissipation is handled also varies between the
methods, Two main classes of procedure can be identified — those that use inelastic
spectra, and those that use equivalent viscous damping. Inclastic spectra are generally
related to acceleration, though there is no inherent reason why inelastic displacement
spectra cannot be generated (sce Section 3.4.3(¢)). They are generated by single-degree-of-
freedom analyses of structures of different initial elastic periods, using a specified
hysteresis rule, and 2 specified maximum ductility. Since the ductility demand cannot
generally be predicted prior to the analyses, the analyses are carried out using a range of
specified force-reduction factors, and the spectrum for a given ductility factor is found by
interpolation within the results of the analyses. Alternatively, simplified relationships
between force-reduction factor and ductility that vary between equal-displacement at long
periods, and equal energy at short periods are directly generated. An example based on
this approach, using the basic EC8™ acceleration spectrum for firm ground and peak
ground acceleration of 0.4g is shown in Fig.1.20(a).
As will be apparent from the discussion related to Fig. 1.18, different inelastic spectra
would need to be generated for different structural systems and materials that exhibited
different hysteretic characteristics. Methods for generation of inelastic spectra are
discussed in Section 3.4.3(e).
The second alternative is to represent ductility and energy dissipation capacity as
equivalent viscous damping, using relationships based on inelastic time-history analyses.
This procedure is only appropriate when the secant stiffness to maximum response is
ased in the design process. The procedure for design using displacement spectra requires
little or no iteration and hence is termed Direct Dighlacement-based Seismic Design (DDBD).34 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
‘The method is discussed in detail in Chapter 3.
Aa example of a spectral displacement set for different damping levels is shown for
the displacement spectrum of EC8, firm ground, 0.4g PGA, in Fig.1.20(b). It will be
evident chat a single spectral set, covering the expected tange of equivalent viscous
damping, will apply for all hysteretic characteristics, provided the relationships between
equivalent viscous damping, ductility, and hysteretic rule have been pre-calibrated by
inelastic time-history analyses. It is also possible, as discussed above, to express the
displacement spectra in terms of ductility, rather than equivalent viscous damping, in a
form analogous to that used for the acceleration spectra of Fig.1.20(a). It will be shows in
Chapter 3 thac inclastic displacement spectra can be generated using precisely the same
data and analyses used to gencrate the rvles relating ductility to damping for a given
hysteresis rule, and that the approaches are then directly equivalent. The disadvantage of
this approach is that inelastic spectra must be generated for each hysteresis rule, and the
determination of equivalent system ductility requires careful consideration
%
O5
z =
< 2 04
g g
= §
Z g 03
< a o2
i Boo
a 2
0
0 1 2 3 4 5
Period T (seconds) Period T (seconds)
(a) Inelastic Acceleration Spectra (b) Damped Displacement Spectra
Fig.1.20 Alternative Spectral Representations of Ductility for EC8 Firm Ground,
PGA=0.4g.
1.4.4. Choice of Design Approach
Comprehensive presentation and comparison of different displacement-based designs
methods is available in two recent documents". 4, Apparent in these and other recent
documeats is a plethora of different nomenclature to describe the new design processes.
This includes the use of “Displacement-Based Design”, “Limit.states Design", Performance Based
Design” and “Consequence Based Design” amongst others. In our view, all attempt generally
the same goal: thar of providing satisfactory displacement solutions to seismic design
problems, and so the term “Displacenent-Based Design” will be used exclusively in this book.Chapter 1. Introduction: The Need for
splacement-Based Seismic Design 38
It is also our view that the Direct Displacement Based Design approach is the most
intellectually satisfying, and best equipped to address the deficiencies of conventional
force-based design, which were presented in some detail in Section 1.3. This approach
has also been developed in rather more complete form than other methods, and has been
applied to a wider category of structures. Finally, we claim that the method is simpler to
apply, and better suited t0 incorporation in design codes. Because of the simplicity of
generation, and wider applicability, representation of hysteretic energy absorption by
equivalent viscous damping will be preferred to the use of inelastic spectra.
‘As a consequence of these considerations, the theoretical basis of the DDBD
approach is developed in detail in Chapter 3. This is preceded in Chapter 2 by a short
discussion of relevant aspects relating to seismicity and intensity characterisation, while
Chapter 4 covers analytical tools appropriate, or necessary, for Direct Displacement-
Based Design. Chapters 5 to 12 describe application of the method to different types of
structural systems, while Chapter 13 discusses application of the DDBD procedures to
assessment of existing structures. Finally, Chapter 14 presents the design method for
buildings in a typical “Code plus Commentary” format, to be used as a possible format for
future codification2
SEISMIC INPUT FOR DISPLACEMENT-BASED DESIGN
2.1 INTRODUCTION: CHARACTERISTICS OF ACCELEROGRAMS
Our understanding of the response of structures to earthquakes, and our design
methodologies, either force-based or displacement-based, are critically dependent on
recordings of strong ground motion by accelerographs. Accelerograms are recordings of
ground acceleration made by accelerographs during earthquakes, and the earliest records
date back to the 1930's. Early accelerograms were recorded in analogue form on
photographic film, and required digitization to put them in a form where their
characteristics could be examined. Accuracy was limited, and the dynamic characteristics
of the accelerographs themselves meant that useful data for preparing response spectra
could be extracted only up 10 periods of about two to three seconds. In the past twenty
years, digital accelerographs, with much higher resolution and longer range of period
integrity have become increasingly common, and the quality of data from recorded
earthquakes is steadily improving as more digital records become available.
This text will not attempt to present seismological information about source
mechanisms, physical and temporal distributions of carthquakes, atterwation relationships
and modern developments in source modelling. The interested reader is encouraged to
read any of the many specialized seismological texts (e.g. $7, K8]. The treatment here will
be limited to information of specific relevance for displacement-based seismic design.
Nevertheless, a brief treatment of some of the common terms and a similarly brief
discussion of the characteristics of accelerograms is warranted.
‘The vast majority of earthquakes are initiated on or adjacent to tectonic plate
boundaries by the slow relative movement of the plates, These are termed interplate
earthquakes. Intraplate earthquakes, occurring far away from plate boundaries are less
common, but nevertheless can be significant for specific sites (for example the New
Madrid region of the central USA, Charleston, South Carolina, USA and various parts of
Australia),
The two basic terms used to provide a measure of the importance of a particular
earthquake are the magnitude and intensity. The magnitude, normally related to the
Richter scalel?4l is a measure of the energy release at the fault zone, while the intensity is a
measure of the local significance of ground motion at a given site, as described by locally
3738 Priestley, Calvi and Kowalsiy. Digplacement-Based Seismic Design of Structures
recorded accelerograms, or by subjective scales, such as che Modified Mercalli scale. The
magnitude depends on the length and transverse dimension of fault that fractures during
the earthquake, and on the average stress-drop in the rock immediately adjacent co the
fatle, resulting from rupture. A magnitude 5.0 -5.5 earthquake may result from faulting
over a length of a few km, while a magnitude 8 earthquake may involve fault slip over a
length as much as 400 km. The energy release is related to magnitude in proportion to
105, implying that the energy released increases by a factor of 32 for each unit increase
of magnitude, Earthquakes wich low magnitude occur frequently, and those of large
magnitude accur less frequently. Averaged globally the relationship between magnitude
and annual probability of occurrence agrees well with a Gumbel extreme type f
distributionl"!, However, this relationship becomes less reliable as the area sampled
reduces in size. The concept of a stationary value for the annual probability of occurrence
of small to moderate earthquakes near a given site is generally reasonable, but for lergee
earthquakes, particularly where a site is affecced predominantly by earthquakes on a single
fault, this assumption may be less valid. Immediately after a major earthquake and its
related aftershocks have ended, the probability of major fault movement of the same
section of fault is significantly reduced, potentially reducing the major contribution to
Jocal seismic hazard. An example is the stretch of plate boundary along the coast of Chile,
where major carthquakes tend to occur on specific segments of the fault at rather regular
time intervals, and with comparatively uniform magnitudes. This of course will not be the
case with smaifer earthquakes, and even for large earthquakes where fracture of one
segment of a fault may create additional stress on the adjacent section, increasing the
probability of fracture of this section in the near future. An example is the Anatolian fault
in Turkey, and the subduction boundary between the Nazca and South American tectonic
plates where fault rupture tends to occur on successive adjacent segments of the fault in a
comparatively tegular sequence. Nevertheless, it is common in seismic hazard analysis to
assume that the Jocal isk is time-invariant
Intensity is dependent on magnitude of the causative earthquake, distance from the
fault zone, mechanism and direction of rupture propagation, and ground conditions at
the site at which intensity is observed, and between the fault zone and the site. There is,
‘no exact means of measuring intensity, since it is generally assessed through the effect
that the earthquake has at a given site on the built environment. This has typically been
defined in the past through descriptive scales such as the modified Mercalli: scale")
Attempts to relate such scales that are dependent on observations of damage to different
structural types and materials, to such measurable quantities as peak ground accelesation
or velocity have not been particularly successful, as different ground motion quantities
have different significance to different structural types. Thus peak ground acceleration
may be important to structures that have brittle failure modes, but may be of litte
importance to a flexible well-confined structure. Duration of shaking may be a key
parameter for a flexible structure without adequate confinement
No two accelerogeams are identical, even when the earthquakes originate in the same
part of a fault, with similar magnitudes, and the site where the accelerograms are recorded
is the same. Some of the differences and similarities between accelerograms are illustratedChapter 2. Seismic Input for Displacement-Based Design 39
by the three examples of Fig.2.1. The first of these is from the moderate Whittier
earthquake of 1987 (Mw = 6.0), recorded in analogue form at a distance of 15km from
the rupture, The second is from the Mw = 6.7 Northridge earthquake of 1994, recorded
at the Sylmar site, at a distance of 6km ftom the fault rupture, and the final record is
from the Mw = 6.9 1995 Kobe earthquake, recorded immediately adjacent to the faul.
None of the records shows the complete duration of recorded motion, but all include the
section of greatest interest, including before, during and immediately after the strong
ground motion. All three records are plotted to the same time and acceleration scales.
The record from the smaller Whittier earthquake has a comparatively short period of
strong ground motion compared with the other two records, the peak ground
acceleration (PGA) is lower, and it appears that high frequency components are more
dominant, On the other hand, all records show an initial period of comparatively high
frequency/low amplitude acceleration before the onset of the strong-motion period of
response, corresponding to the time period between acrival of the P and $ waves. Both
the Sylmar and Kobe records show high amplitude/low frequency pulses in the initial
stages of the strong ground motion, corresponding to a focusing effect related to the
mechanism of energy release and the local geology, known as a velocity pulse. In the case
of the Kobe record, this is primarily a result of forward directivity where the fault
fractures over a short period of time from one end to another, focusing the energy in the
downstream direction, In the Sylmar record, the reasons are apparently more complex,
involving basin edge effects!.
Ir is also of some interest to examine the time sequence of ground displacement,
found by double integration of the acceleration records. These are shown for the same
three records in Fig.2.2, and are plotted to the same time scales, but with a factor of 30
difference between the displacement scales of the Whittier and the other two records. It
should be noted that integration of the acceleration records to obtain displacement
records is inevitably subject to some error. Small systematic errors in the acceleration
record can lead to large errors in the displacement record, causing the apparent
displacement to drift in one direction. Base-line corrections are typically carried out to
remove this drift, but the accuracy of such corrections is uncertain,
It will be apparent from comparison of Figs.2.1 and 2.2 that the differences between
the displacement records are more pronounced than between the acceleration records.
Although the PGAs for the three records only vary by a factor of about 2.5, the peak
ground displacements (PGD) vary by a factor of about 25. All three records exhibit much
less high frequency content in displacement terms than in acceleration terms, but the
Whittier record is significantly richer in high frequency components than the other nwo
records. The two more intense records appear to show dominant long-period ground
displacement response. In the case of the Sylmar record, this appears to correspond to a
period of about 3 to 4 seconds.
Intensity, for a given earthquake, decreases with distance from the fault, Attenzacion
relationships are used to describe this reduction in intensity. However, there is a large
spatial variation in recorded ground motions between different sites at equal distances
from the epicentre of an earthquake, Attenuation relationships are averages found fromPriestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
085
oa
Acceleration (xg)
;
S
L
(2) Whittier Earthquake, My=6.0, 1987
08 pg
0 4 8 2 6 2
Time (see)
085
044
Acceleration (x8)
i
oa
(©) Sylmar Record, Northridge Earthquake, My=6.7, 1994
08 J 4
o 4 8 2 6 2
Time (see)
085
Aceleraton() _
joi
(©) Kobe Earthquake, M
084
° 4 8 2 16 20
“Time (sec)
Fig.2.1 Selected Time-Windows of Different AccelerogramsChapter 2, Seismic Input for Displacement-Based Design 4
(8) Whittier Record My6.0, 1987
10 5
Time (see)
Displacement (em)
oo
(b) Sylmar Record, 1994 Northridge Earthquake, My
“30 St oo?
o 5 0 5 2»
Time (see)
30
24
Displacement (em)
L
earthquake 1995, My=6.9
|
o 5 ny 5 ™
Fig.2.2 Selected Time-Windows of Ground Displacement ftom the
Accelerograms of Fig.2.142 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
recorded accelerograms, and typically do not account for many of the factors known to
influence intensity. Somerville et al’; have noted that the variations in ground motion,
which are particularly apparent at periods greater than one second (and hence are of
particular importance to structural design — whether force-based or displacement-based)
“can usually be attributed to features of the earthquake source such as the orientation of
the fault plane, the style of faulting (strike-slip or dip-slip), and the evolution and
distribution of slip on the faule plane”, Thus it is reasonable to assume that the current
uncertainty associated with site intensity predicted by probabilistic seismic hazard analyses
(PSHA) will be reduced as site and source modelling improves.
A number of the factors affecting spatial variaion of ground motion from a given
earthquake are discussed in [S5] in relation to the Los Angeles Basin, The folowing notes
provide 2 summary of the discussion in [S5]:
Near fault rupture directivity pulse: Near fault recordings from recent earthquakes
indicate that ground motion is dominated by a large long-period narrow-band pulse in the
fault-normal motion, whose petiod increases with magnitude! This pulse may have a
dominant period of about 1 sec. for earthquakes of magnitude My = 6.7 - 7.0, and as high
as 4 sec. for earthquakes of magnitude My = 7.2 - 7.6
Reverse faulting earthquakes: Ground motions from reverse faulting earthquakes
are systematically stronger than ground motions from strike-slip earthquakes. The
influence may be as much as 20-40%,
Buried faulting earthquakes: Ground motions from shallow earthquakes that do
not break the ground surface are systematically stronger than from earthquakes that result
in surface faulting. Again the influence may be in the order of 20 40%, The 1989 Loma
Prieta, and the 1994 Northridge earthquakes are examples of shallow earthquakes without
surface faulting.
Ground motion from lange surtace faulting earthquakes: Ground motions from
earthquakes that produce large surface faulting (e.g, the Chi-Chi earthquake in Taiwan)
tend to be significantly lower than predicted by current ground motion models, and
substantially lower than ground motions from buried faulting earthquakes.
Basin effects: Current codes modify design ground motions on the basis of the
shear-wave velocity in the upper 30m of soil. This is only appropriate for rather short-
period motion, as at periods greater than one second, seismic wave lengths are much
longer than 30m, and response is likely to be influenced by soil properties at depths of
hundreds, and perhaps thousands of metres. Basin edge effects can also be significant,
with constructive interference between waves entering from the edge and from the basin
below, particularly when the basin has steep fault-controlled margins.
Recently developed hybrid simulation procedures‘! are capable of incorporating all
of the above features in calculating broadband ground motion time-histories for
prescribed earthquake scenarios. PSHAs based on these techniques are already more
reliable than those based on attenuation relationships, and it can confidently be expected
that improved characterization of seismicity of specific sites will continue. Ir is likely chat
future developments will be less towards improved accuracy of code spectra, thanChapter 2. Seismic Input for Displacement-Based Design 43
towards improved mapping of local characteristics defining seismicity, with spectral shape
as well as spectral ordinates being a mapped variable.
2.2. RESPONSE SPECTRA
2.2.1 Response Spectra from Accelerograms
‘The fundamental information extracted from accelerograms or PSHA’s for use in
design is typically expressed in the form of response spectra, which represent the peak
response of single-degree-of-freedom oscillators of different periods of vibration to the
accelerogram. The quantities most commonly represented in response spectra are
absolute acceleration (with respect to “at rest” conditions), and relative displacement
(with respect to instantaneous ground displacement), though relative velocity response
spectra are also sometimes computed. The procedure is represented in Fig.2.3 where five
different SDOF oscillators are depicted in Fig.2.3(a) subjected to the earthquake ground
motion ag, ‘The peak absolute acceleration and relative displacements recorded during
response to the accelerograms are plotted against the period of the structure in Fig.2.3(b).
Normally response spectra provide information on the peak elastic response for a
specified elastic damping ratio (typically 5%), and are plotted against the elastic period. It
is, however, also possible to plot inelastic spectra related to specified displacement
ductility levels. In this case the period may represent the initial elastic period, or the
efiective period at peak displacement demand, related to the effective stiffness
acceleration
acceleration
displacement
TT T2) 13) T4 OTS
Period
(a) SDOF Oscillators (b) Elastic Response Spectra
Fig.2.3 Formation of Response Spectra
Examples of elastic acceleration and displacement spectra for the chree accelerograms
represented in Figs.2.1 and 2.2 are shows ia Fig,2.4, The spectra are shown for four levels
of elastic damping, expressed as ratios to the critical damping. Some interesting
conclusions can be drawa from examination of these figures. The Whittier accelerogram
has a PGA of about 0.4g (see Fig.2.1(a)), and a peak response acceleration of more than44 Priesdey, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
085 —
3 é =0.05,
ei. 2 16 0.10
g 06 3 01s
a & E12 020
yo =
g Fos
Boa z
an 5044
& B
° T TT 1 ° T TTT 4
o 12 3 5 o 203 4 8
Period (seconds)
Spectral acceleration (xg)
Period (seconds)
(a) Whittier (My=6.0)
8
3
L
Spectral displacement (cm)
8 &
L 1
Period (seconds)
Lit
Spegtral acceleration (xg)
(b) Sylmar (Northri
8 $s 8 8
botiii
Spectral displacement (cm)
i
2 3
Period (seconds)
idge,1994; My =6.7)
4005
3
Period (secands)
(©) Kobe (My=
9)
Period (seconds)
Fig.2.4 Acceleration and Displacement Elastic Response Spectra for
Accelerograms of Figs.2.1 and 2.2Chapter 2. Seismic Input for Displacement-Based Design 45
0.8g at a period of about 0.25 sec. This might be considered to represent reasonably
strong ground motion, since design PGA’s and peak response accelerations for high-
seismicity regions are often in the range 0.4g and 1.g respectively. However, when we
examine the displacement spectra from the same accelerogram, we find that the peak
response displacement is less than 20mm (0.79in}, for a damping level of 5% of critical
damping, Thus, if 2 given structure is capable of sustaining this very minor peak response
displacement within the elastic range of response, no damage would be expected, despite
the high peak response acceleration. To put this in perspective, information provided in
Section 4.4.6 indicates that for reinforced concrete frame buildings of typical proportions,
effective yield displacements for two- and four-storey buildings might be approximately
45mm (1.8 ia) and 90mm (3.6in) respectively — significantly larger than the peak
displacement response for this accelerogram, It is apparent that only very stiff and brittle
structures would be expected ¢o be at risk from an accelerograms similar to the Whittier
record. This is in agreement with the recorded damage in the Whittier earthquake.
Another point of interest is apparent from Fig. 24(a). Information from the
acceleration response spectra cannot be extracted for periods of T > 1.5 sec since the
response accelerations are so low. The displacement spectre provide much mare readily
accessible information for the medium to long period range, but indicate surprisingly
regular displacements at periods greater than about 2 seconds. In fact this is false daca,
since the accelerogram was recorded by an analogue, rather than digital accelerograph,
and a filter at 3 seconds was used to determine the displacement response. Bommer et
all have shown that the roll-off associated with filtering makes the response spectra
unreliable for periods greater than about 2/3rds of the filter period. Thus the data in the
displacement spectra of Fig,2.4(a) are meaningless for periods greater chan about 2 sec
The Northridge Sylmar acceleration spectra of Fig.2.4(b) show peak acceleration
response for 5% damping of about 2.7g — about three times the response for the Whictier
earthquake. The displacement spectra, which resuit from a digital accelerograph, and are
reliable up to significantly longer periods, indicate peak displacement response of about
800mm (31.5in) ~ more than 40 times that of the Whittier record, Clearly this record
would be expected to have much greater potential for damage than the Whittier record
Note that after reaching a peak response at about 3 sec., displacement response decreases
at higher periods.
The Kobe record of Fig.2.4(0) also has high peak displacement response, and
somewhat similar characteristics to the Sylmar record, though the peak displacement
response appears to occur at a reduced period,
2.2.2 Design Elastic Spectra
(a) Elastic Acceleration Spectra: Until recenuy, design spectra for seismic design of
seructures were typically specified in design codes as a spectral shape related t0 soil
conditions, modified by a design PGA, reflecting the assessed seismicity of the region
where the structure was to be built. Typically only acceleration spectra were provided, and46 Priestley, Cah and Kowaisky. Displacement-Based Seismic Design of Structures
mapping of the variation of PGA with location was coarse. This is still the case with
many seismic design codes.
Recently, more detailed information has been provided in different design codes, such
asathe IBCI4 of the USA, and the new Italian scismic design codel*", where spectral
acceleration ordinaces at two or three key periods are provided for a given site for
different return periods. Typically this is provided through a computerized data base,
enabling design data to be extracted based on site longitude and latitude. However, this is
typically provided only for acceleration response spectra; peak displacement response is
not yet available in many design codes.
“The typical form of elastic acceleration response spectra is illustrated in Fig.2.5(a). The
shape is smoothed, reflecting the average of many accelerograms, and is based on
probabilistic estimates of the contribution to seismic risk of a larger number of smaller
earthquakes, and a reduced number of larger earthquakes. The result is a spectrum where
the acceleration ordinates have uniform probability of occurrence for a given return
period (ee Section 2.2.2(c) below).
The spectrum tises from the PGA at T= 0 to a maximum value ata period Ty
(typically about 0.15 seconds). For soft soils, codes typically amplify the PGA above the
value applicable for firm ground, or rock. ‘The plateau typically has a response
acceleration of about 2.5 to 2.75 times the PGA, The acceleration plateau continues to a
period of Tp, the value of which depends on the ground conditions in the near-surface
layers, with larger values applying to soft soils, as indicated in Fig;2.5(a). The value of Ty
also typically depends on the magnitude of the earthquake, as is apparent from Fig.2.4,
with smaller values being appropriate for earthquakes of lower magnitude, For periods
greater than Tp the response acceleration reduces, typically in proportion to T, implying 2
constant-velocity response. In many codes this constant-velocity part of the spectrum
continues indefinitely. More advanced codes specify an upper limit of T= Te for the
constant-velocicy range, above which the acceleration decreases in proportion to 7. A
completely opposite trend is apparent in some less advanced codes, where a constant
plateau corresponding to a minimum specified response acceleraion is sometimes
defined. This is shown by che dash-dot line in Fig,2.5(a). The intent of such a provision is
to ensure that the lateral strength of a structure is nor less than a code-specified minimum
value. However, this is better controlled by limits on P-A moments (see Section 3.8). As
discussed below, when che logic of the minimum acceleration plateau of the acceleration
spectrum is translated to equivalent displacements, impossible trends result.
“The general form of the clastic acceleration spectrum can be defined by che following
equations:
0 2000 Mans
e
é
J 1600
5 1200
= 800
a Mono
= 400
Z M=65
& M=60
o 2 4 6 8 w
Period (seconds)
(@) r= 10 km
rey
1600 4
§ 1200 4
g Mens
soo |
404 a0
MEGS
o 2 4 6 8
Period (seconds)
(0) c= 20 kon
52000),
T1600 5
§ 1200 4
z
= soo
a M=7s
= 04
a
& 4 M
o 2 4 6 8 w
Period (seconds)
(r= 40 km
Fig.2.7 Influence of Magnitude and Distance on 5% Damped Displacement
Spectra for Firm Ground Using Eqs.(2.3) and (2.5) [after F6]
5152 Priestley, Calvi ancl Kowalsky. Displacement-Based Seismic Design of Structures
based design in a form similar to the maps of spectral accelerations at key defined periods
seems likely to end up with mapped corner periods and peak response displacements in a
data base related to GPS coordinates. Preliminary results from this study indicate that
both the corner period and the peak elastic response period given by Eqs.(2.3) and (2.5)
respectively may be revised upwards by approximately 20%.
Figure 2.8 compares the different equations for corner period as a function of
magnitude. It is clear that the current EC8 equation is severely non-conservative, and
there are significant differences berween the NEHRP equation, and the
equation determined from the work by Faccioli et alli, It appears that the work in
progress in Italy, using a world-wide data-base of some 1700 digital records, and shown
tentatively in Fig2.8 by chree dots interpreted from data supplied by Paccioli® may lie
somewhere between Eq,(2.3) and (2.6)
Figure 2.9 shows average displacement spectra interpreted by Faccioli et al from the
world-wide data base of 1700 digital records for earthquakes of magnitude 6.4>~-~ £4.29)
gos Tees)
* 024
»f—.————-. ,
0 01 02, 030d
Damping Ratio
Fig.2.12 Damping Modifiers to Elastic Spectral Displacements60 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
ce)
| |
Spectral Displacement (em)
° 1 2 3 4
Period (seconds)
Fig. 2.13 Comparison of Spectral Analysis Results and Eq.(2.8) for a Suite of 14
Spectrum-Compatible Accelerograms
might be appropriate. Ic will be seen that Eg.(2.11) is the same as Eq.(2.8) but with @=
0.25 instead of 0.50. The effect of this modification is to increase the value of Rg
compared with the value applying for “normal” accclerograms. The seduction factor
resulting from Eq.(2.11) is induded ia Fig.212 for comparison with the other
expressions, and Fig.2.14 compares the dimensionless displacement modifiers for the
1998 EC8 expression (Eq.2.8) and the expression suggested for near-field forward
directivity conditions (Ig,(2.11)). ‘The data in Fig.2.14 are based on the shape of the
displacement spectrum for firm ground plotted in Fig. 2.5(a). Some qualified support for
5g.2.11) is available in work by Bommer and Mendis who provide additional
discussion of this topic. Their work indicates that the scaling factors may be period-
dependent, which is not currently considered in design
It will be shown in Section 3.4.6 that use of displacement spectra for near-field
forward-directivity effects results in a requirement for higher base-shear strength when
compared t0 requirements for “normal” conditions. This requirement has been
recognized for a number of years, in particular since the 1994 Northridge earthquake, and
is incorporated empirically in recent force based codes [e.g. X4, X8}. This is an cxample
of conditions where it has been recognized that existing displacement-equivalence rules
are inadequate in force-based design. With displacement-based design, the influence of
near-field effects are directly incorporated in the design, provided the reduced influence
of damping (and ductility) in modifying displacement response is recognized by graphs
such as Fig.2.14(b),Chapter 2. Seismic Input for Displacement-Based Design 61
i
e
S
®
°
a
©
a
°
5
°
5
([email protected]))
°
°
nN
(Eq.(2.11))
Relative Displacement A,./A,.,
Relative Displacement A,./A,.,
a 0 a
o 4 2 3 4 5 o 4.2 3 4
Period T (seconds) Period T (seconds)
(a) "Normal" Conditions (b) Velocity Pulse Conditions
Fig.2.14 ECS (1998) Damping Reduction Factor, and Tentative Factor for
Forward Directivity Effects
As discussed earlier in this chapter, displacement spectra for design are in a
developmental stage, with rather rapid progress being made by scismologists. The same
statement can be applied to developments in definition of acceleration spectra. It is
probable that displacement spectra, and modifiers for damping, period and ductility will
change from the tentative values suggested in this Chapter. However, this should not be
taken to reduce the utility of the work presented in subsequent chapters. The approaches
developed are independent of the displacement spectra, and several different shapes will
be used in different examples, to illustrate the flexibility of the direct displacement-based
design method.
It should also be noted that the material presented in this chapter is at least as valid as
acceleration spectra currently used for force-based design (see p53), Taken together with
the extensive research described in Chapters 3 and 4 on relationships between ductility
and damping, the procedures developed are significantly less susceptible to errors than
resulting from current force-based design, and much better adapted to achieving specified
limit states
2.3 CHOICE OF ACCELEROGRAMS FOR TIME-HISTORY ANALYSIS,
The most reliable method at present for determining, or verifying the response of a
designed structure to the design level of intensity is by use of non-linear time-history
analysis. The selection and characteristics of accelerograms to be used for this requires
careful consideration, The reader is referred to Section 4.9.2(h) where this is discussed in
some detail.3
DIRECT DISPLACEMENT-BASED DESIGN:
FUNDAMENTAL CONSIDERATIONS
3.1. INTRODUCTION
‘The design procedure known as Direct Displacement-Based Design (DDBD) has
been developed over the past ten yearsiP8”7I0") wich the aim of mitigating the
deficiencies in current force-based design, discussed in some detail in Chapter 1. ‘The
fundamental difference from force-based design is that DDBD characterizes the
structure to be designed by a single-degree-of-freedom (SDOF) representation of
performance at peak displacement response, rather than by its initial elastic
characteristics. ‘This is based on the Substitue Structure approach pioneered by others\'S2
The fundamental philosophy behind the design approach is to design a structure
which would achieve, rather than be bounded by, a given performance limit state under a
given seismic intensity. This would result in essentially uniform-risk structures, which is
philosophically compatible with the uniform-risk seismic spectra incosporated in design
codes. The design procedure determines the strength required at designated plastic hinge
locations to achieve the design aims in terms of defined displacement objectives. It must
then be combined with capacity design procedures to ensure that plastic hinges occur
only where intended, and that non-ductile modes of inelastic deformation do not
developl”!, These capacity design procedures must be calibrated 0 the displacement-
based design approach. ‘This is discussed further in general terms in Sections 3.9 and 4.5,
and in specific structure-related terms in the appropriate structural chapters. It will be
shown that capacity design requirements are generally less onerous than those for force-
based designs, resulting in more economical structures.
‘This chapter deals with fundamental aspects of the approach that are common to all
materials and structural systems. Subsequent chapters deal with detailed application to
different structural systems, including verification by design/analysis examples.
3.2. BASIC FORMULATION OF THE METHOD,
‘The design method is illustrated with reference to Fig.3.1, which considers a SDOF
representation of a frame building (Fig-3.1(a)), though the basic fundamentals apply to all
structural types, The bi-linear envelope of the lateral force-displacement response of the
SDOF representation is shown in Fig.3.1(b). An initial elastic stiffness Kj is followed by a
6364 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
post yield stiffness of Kj
While force-based seismic design characterizes a structure in terms of elastic, pre-yield,
properties (initial stiffness Kj, clastic damping), DDBD characterizes the structure by
secant stiffness K, at maximum displacement Ag (Fig3.1()), and a level of equivalent
viscous damping representative of the combined elastic damping and the hysteretic
energy absorbed during inelastic response. Thus, as shown in Fig. 3.1(0), for a given level
of ductility demand, a structural steel frame building with compact members will be
assigned a higher level of equivalent viscous damping than a reinforced concrete bridge
designed for the same level of ductility demand, as a consequence of “fatter” hysteresis
loops (see Fig.
(a) SDOF Simulation
028 5 05
sap 2 od
S £
0.15 503
2 g
Boa 3
5 e
a a
Displacement Duetility Period (seconds)
(©) Equivalent damping vs. ductility (d) Design Displacemene Spectra
Fig. 3.1 Fundamentals of Direct Displacement-Based DesignChapter 3. Direct Displacement-Based Design: Fundamental Considerations 6
r
~ 3
(a) Idealized Steel (b) Reinforced Concrete _(c) Friction Slider
Frame Response Frame Response Response
Fores ro F
8
(d) Bridge Column with (e) Non-linear Elastic _() Unsymmetrical
High Axial Load with P-A Strength
Fig.3.2 Common Structural Force-Displacement Hysteresis Response Shapes
Wich the design displacement at maximum response determined, as discussed in
Section 3.4.1, and the corresponding damping estimated from the expected ductility
demand (Section 3.4.3), the effective period J; at maximum displacement response,
measured at the effective height He (Fig.3.1(a)) can be read froma set of displacement
spectra for different levels of damping, as shown in the example of Fig.3.1(d). The
effective stiffness K, of the equivalent SDOF system at maximum displacement can be
found by inverting the normal equation for the period of a SDOF oscillator, given by Eq,
(1.6), to provide
K,=4n'm,/T? Gl)
where me is the effective mass of the structure participating in the fundamental mode of
vibration (see Section 3.5.3). From Fig.3.1(b), the design lateral force, which is also the
design base shear force is thus
2)66 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
‘The design concept is thus very simple. Such complexity that exists relates to
determination of the “substitute structure” characteristics, the determination of the
design displacement, and development of design displacement spectra. Careful
consideration is however also necessary for the distribution of the design base shear force
Vpase to the different discretized mass locations, and for the analysis of the structure
under the distributed seismic force. These will be discussed later.
The formulation of DDBD described above with reference to Figs 3.1(¢) and (d) has
the merit of characterizing the effects of ductility on seismic demand in a way that is
independent of the hysteretic characteristics, since the damping/ductility relationships are
separately generated for different hysteretic rules. It is comparatively straightforward to
generate the influence of different levels of damping on the displacement response
spectra, (see Section 2.2.3) and hence figures similar to Fig. 3.1(d) can be generated for
new seismic intensities, or new site-specific seismicity using standard techniques!6)
It is also possible, however, to combine the damping/ductility relationship for a
specific hysteresis rule with the seismic displacement spectral demand in a single inelastic
displacement spectra set, where the different curves directly relate to displacement
ductility demand, as illustrated in the example of Fig.3.3
Disp.
splacement (mm),
Bo
©
Oo 1 2 3 4 5
Period (seconds)
Fig.3.3. Example of an Inelastic Displacement Spectra Set Related to Effective
Period for a Specific Hysteresis Rule
With the seismic demand characterized in this fashion, the design procedure is slightly
simplified, as one step in the process is removed. The inelastic displacement spectra set is
entered with the design displacement (to be discussed subsequently) and the design
effective period is read off for the level of the design displacement ductility. Although
this is a slightly simplified procedure, it requires that inclastic displacement spectra be
generated for different hysteresis rules for each new seismic intensity considered. Since
this is a rather lengthy process, we will use the formulation of Fig.3.1 in the examples of68 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
© Level 3: Life safe. Life safety is essentially protected, damage is moderate to
extensive
* Level 4: Near collapse. Life safety is at risk, damage is severe, structural
collapse is prevented,
‘The relationship berween the four levels of seismic excitation, and the annual
probabilities of exceedence of each level will differ according to local seismicity and
structural importance, as discussed in Section 2.2.2(c). In California, the following levels
are defined th
© EQ: 87% probability in 50 years: 33% of EQ HI
© EQ: 50% probability in 50years: 50% of EQUI
© EQ-IE: approximately 10% probability in 50 years.
© EQ-IV: approximately 2% probability in 50 years: 150% of EQUI,
‘The relationship between these performance levels and earthquake design levels is
summarized in Fig3.4. In Fig.34 the line “Basic Objective” identifies a series of
performance levels for normal structures. The lines “Essential Objective” and “Safety
Critical Objective” relate performance level to seismic intensity for nwo structural classes
of increased importance, such as lifeline structures, and hospitals. As is seen in Fig.3.4,
with “Safety Critical Objective”, operation performance must be maintained even under
the EQUIV level of seismicity
Although the Vision 2000 approach is useful conceptually, it can be argued that it
requires some modification, and that it provides an incomplete description of
performance. The performance levels do not include a “damage control” performance
level, which is clearly of economic importance. For example, it has been noted that
although the performance in the 1995 Kobe earthquake of reinforced concrete frame
‘System Performance Level
Fully Near
Operationat OPerational Life Safe
Frequent | O [ae
(43 year) Unacceptable
Performance 4
Occasional oe ew contraction) 2
(72 year)
|
Rare
(475 year)
Very Rare
(970 year)
Fig. 3.4 Relationship between Earthquake Design Level and Performance Level
(after Vision 2000 |02!)Chapter 3. Direct Displacement-Based Design: Fundamental Considerations 69
buildings designed in accordance with the weak-beam/strong-column philosophy
satisfied the “Life safe” performance level, the cost of repairing the many locations of
inelastic action, and hence of localized damage, was often excessive, and uneconomic!),
‘The performance level implicit in most current seismic design codes is, in fact, a damage
control performance level. Further, it has been argued that residual drift, which is not
considered in the Vision 2000 classification, should be considered an important measure
of performance!”
In order to better understand the relationship between structural response levels and
performance levels, it is instructive to consider the relationship between member and
structure limit states!
3.3.1. Section Limit States
(a) Cracking limit state: For concrete and masonry members the onset of cracking
generally marks the point for a significant change in stiffness, as shown in the typical
moment-curvature relationship of Fig,3.5(a). For critical members expected to respond in
the inelastic range to the design-level earthquake, this limit state has little significance, as
itis likely to be exceeded in minor seismic excitation, even lower than that corresponding,
to the EQ level of Vision 2000. The limit state may, however, be important for
members that are expected to respond essentially elastically to the design-level
earthquake. For example, the appropriate stiffness to be used for a prestressed bridge
superstructure will depend on whether or not the cracking limit state is exceeded.
(Q) First-yield limit state: second significant change in stiffness of concrete and
masonry members occurs at the onset of yield ia the extreme tension reinforcement.
‘This is also the case in structural steel members. ‘This limit state is useful for defining the
appropriate clastic stiffness to be used in analyses of duciile systems using simplified
hysteresis rules, such as biclinear response, shown in Fig.3.5(a) by the dashed line.
(c) Spalling limit state: With concrete or masonry sections, the onset of spalling of the
cover concrete or masonry may be a significant limit state, particularly for unconfined
sections, or sections subjected to high levels of axial compression, where spalling is
typically associated with onset of negative incremental stiffness and possibly sudden
strength loss. Exceedence of this limit state represents a local condition that can be
expected to require remedial action. For well-confined sections, this is likely to be the
only significance of onset of spalling, since the member can be expected to support much
larger deformations without excessive distress. Strength may in fact continue to increase
beyond this limit state, Conservatively, a compression strain of € = 0.004 may be
assumed for concrete structures, and a rather lower value for masonry structures
(d) Buckling limit state: With reinforced concrete or masonry members, initiation of
buckling of longitudinal reinforcement is a significant limit state. Beyond this limit state,
remedial action will often require removal and replacement of the member. With70 Priestley, Calvi and Kowalsky. Displacement-Based Scismic Design of Structures
structural steel members, particularly flanged beams and columns, onsct of buckling also
represents a significant limit state for the same reasons as for concrete members.
Spalling
Moment
Force
Yield
Cracking
\
1
i
1
|
|
Ay A
a, A
a Au
Curvature Displacement
(a) Section Limit States (b) Structure Limit State
Fig.3.5 Member and Structure Design Limit States
(©) Uhimate limie state: Definition of the ultimate limit state for members is
somewhat subjective. It is sometimes taken to correspond to a critical physical event,
such as fracture of confinement reinforcement in a potential plastic hinge zone of a
concrete member, or weld fracture of a structural steel connection. Another common
definition relates ¢o a specified strength drop (20% is often used) from the maximum.
attained (or sometimes from the design) strength. Neither definition truly corresponds to
an ultimate limit state, since at least some residual strength is maintained for further
increase of displacement. A true ultimate limit state would refer to inability to carry
imposed loads, such as gravity loads on a beam, or axial forces in a column. However, the
occurrence of negative incremental stiffness of the moment-curvature characteristic,
which is associated with strength drop, is cause for concern under dynamic response,
since it implies redistribution of strain energy from elastically responding portions of the
steuctuce into che member with negative stiffness. This has potentially explosive
consequences.
3.3.2 Structure Limit States.
(a) Serviceability Limit State: This corresponds to the “fully functional” seismic
performance level of Vision 2000, No significant remedial action should be needed for a
structure that responds at this limit state. With concrete and masonry structures, no
spalling of cover concrete should occur, and though yiekd of reiaforcemeat should be
acceptable at this limit state, residual crack widths should be sufficiently small so that
injection grouting is not needed. As suggested by Fig.3.5(b), structural displacements at
the serviceability limit state will generally exceed the nominal yield displacement.
For masonry and concrete structures this limie stare can be directly related co strain
limits in the extreme compression fibres of the concrete or masonry, and in the extremeChapter 3. Direct Displacement-Based Design: Fundamental Considerations 1
tension reinforcement, With structural stcel buildings, the limit state is more likely to be
related to non-structural elements, as discussed in the following.
Potential for non-structural damage must also be considered when determining
whether or not the serviceability limic state has been exceeded. Ideally, non-structural
elements, such as partition walls, and glazing, should be designed so that no damage will
occuz to them before the structure achieves the strain limits corresponding ro the
serviceability limit state. Even with brittle partitions this can be achieved, by suitable
detailing of the contact between them and the structure, normally involving the use of
flexible jointing compounds. However, when the typical construction involves brittle
lightweight masonry partitions built hard-up against the structure, significant damage to
the partitions is likely at much lower displacement levels than would apply to the
structure, For example, reinforced concrete or structural steel building frames are likely to
be able to sustain drifts (lateral displacements divided by height) of more than 0.012
before sustaining damage requiring repair. In such cases, the serviceability limit state is
unlikely to govern design. A very different conclusion will result if low-strength
lightweight masonry infill is placed in the frames, without flexible connection. The infill i
likely to reach its limit state at drift levels less than 0,005, and design to avoid non-
structural damage in the infill may well govern the structural design. This is considered in
more detail in Chapter 5.
(6) Damage-Control Limit State: As noted above, this is not directly addressed in the
Vision 2000 document, but is the basis for most current seismic design strategies. At this
limit state, a certain amount of repairable damage is acceptable, but the cost should be
significantly less than the cost of replacement. Damage to concrete buildings and bridges
may include spalling of cover concrete requiring cover replacement, and the formation of
wide residual flexural cracks requiring injection grouting to avoid later corrosion,
Fracture of transverse or longitudinal reinforcement, or buckling of longitudinal
reinforcement should not occur, and the core concrete in plastic hinge regions should not
need replacement, With structural steel buildings, flange or shear panel buckling should
not occur, and residual drifts, which tend to be larger for structural steel than concrete
buildings should not be excessive. With well designed structures, this limit state normally
corresponds to displacement ductility factors in the range 3 < fly <6.
Again, non-structural limits must be considered to keep damage to an acceptable level
This is particularly important for buildings, where the contents and services are typically
worth three to five times the cost of the structure. It is difficule to avoid excessive damage
when the drift levels exceed about 0.025, and hence it is common for building design
codes to specify drift limits of 0.02 to 0.025. At these levels, most buildings - particularly
frame buildings - will not have reached the structural damage-control limit state. It will be
noted that this limitation will not normally apply to non-building structures such as
bridges and wharves, and consequently structural limits will govern design to the damage-
control limit state for these structures. fective drift limits for these structures are often
in the range 0.03 to 0.045. This limit state is represented in Fig.3.5(b) by the displacement
AaR Priestley, Calvi and Kowalsky. Displacernent-Based Seismic Design of Structures
(c) Survival Limit State: I is important that a reserve of capacity exists above that
corresponding to the damage-control. limit state, to ensure that during the strongest
ground shaking considered feasible for the site, collapse of the structure should not take
place. Protection against loss of life is the prime concern here, and must be accorded high
priority in the overall seismic design philosophy. Extensive damage may have to be
accepted, to the extent that it may not be economically or technically feasible to repair the
structure after the earthquake, In Fig.3.5(b) this limit state is represented by the ultimare
displacement, Ay.
Although the survival limit state is of critical importance, its determination has received
comparatively little attention, Clearly this limit state is exceeded when the structure is no
longer able to support its gravity loads, and collapses. This occurs when the gravity-ioad
capacity is reduced below the level of existing gravity loads as a result of (say) total shear
failure of a cxitical column, resulting in progressive collapse. Alternatively, collapse results
failure, when the P-A moments exceed the residual capacity of the
structure, as illustrated in Fig.3.6 for a bridge column. If the ultimate displacement
capacity assessed from the intersection of the resistance and P-A curves exceeds the
maximum expected in the survival-level earthquake, collapse should not occur.
from a stabili
Strength P-A Moment
Base Moment M
be
Displacement A
Fig.3.6 P-A Collapse of a Bridge under Transverse Response!!!
3.3.3. Selection of Design Limit State
The discussion in the previous sections indicates that a number of different limit states
or performance levels could be considered in design. Generally only one — the damage-
control limit state, or at most owo (with the serviceability limit state as the second) will be
considered, except for exceptional circumstances. Where more than one limit state is
considered, the required strength to satisfy each limit will be determined, and the highest
chosen for the final design. More information on strain and drift limits corresponding toChapter 3. Direct Displacement-Based Design: Fundamental Considerations 3
different performance levels are included in Section 4.2.5, and in the relevant chapters on
different structural systems.
3.4 SINGLE-DEGREE-OF-FREEDOM STRUCTURES
3.4.1 Design Displacement for a SDOF structure
The design displacement will depend on the limit state being considered, and whether
structural oF non-structural considerations are more critical. For any given limit state (see
previous section) structural performance will be governed by limiting material strains,
since damage is strain-related for structural clements. Damage to non-structural elements
can be generally considered drift-related.
It is comparatively straightforward to compute the design displacement from strain
limits. Consider the vertical cantilever structure of Fig.3.7(a). The most realistic structure
conforming to the assumptions of a SDOF approximation is a regular bridge under
wansverse excitation, Two possible reinforced concrete sections, one circular and one
rectangular are shown in Fig.3.7(b). The strain profile at maximum displacement response
is shown together with the sections. Maximum concrete compression strain & and
reinforcement tensile strain & are developed. The limit-state strains are &je and Bye for
concrete compression and steel tension respectively. These will not generally occur
simultaneously in the same section, since the neutral axis depth ¢ is fixed by the
reinforcement ratio, and the axial load on the section. Consequently there are two
possible limit state curvatures, based on the concrete compression and the reinforcement
tension respectively:
Poe = Eg 1C (concrete compression) (33a)
(reinforcement tension) (3.3b)
“The lesser of @se and ys Will govern the structural design. The design displacement cart
now be estimated from the approach in Section 1.3.4(a) as
Ayn =A, +4, =9,(H + Ley) 3+, ~ OL, 64)
where gy, is the lesser of Pee and ys , Ayis the yield displacement ( see Section 3.4.2), Hf
is the column height (see Fig.3.7), Lsp is the effective additional height representing strain
penetration effects (see Section 4,2.7) and Ly is the plastic hinge length.
If the limit state has a code-specified non-structural drift limit @ the displacement
given by Eg.(3.4) must be checked against
yy =O.H 65)74
ley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
(a) Cantilever Bridge Column
TF
Ee
*
be
& &
4 L
(b) Column Sections and Limit State Strains
Fig.3.7 Curvatures Corresponding to Limit Strains for a Bridge Pier
‘The lesser of the displacements given by Eqs. (3.4) and (3.5) is the design displacement.
Note that in many cases the design approach will be to design the structure for a
specified drift, and then determine the details to ensure the strain limits are achieved.
For example, as is shown in Section 4.2.4(a), the limiting concrete strain for the
damage-control limit state is determined from the transverse reinforcement details. Thus
the concrete strain corresponding to the drift limit can be determined by inverting
Egs.(3.4) and (3.3a), and the required amount of transverse reinforcement calculated.
This simplifies the design process.Chapter 3. Direct Displacement-Based Design: Fundamental Considerations 75
3.4.2 Yield Displacement
For a SDOF vertical cantilever, the yield displacement is required for two reasons.
First, if structural considerations define the limit displacement (Eq.3.4), the yield
displacement and yield curvature must be known. Second, in order to calculate the
equivalent viscous damping, the displacement ductility H4y=A4/Ay, which depends on the
yield displacement, must be known
Analytical results presented in Section 4.4 indicate that for reinforced concrete (and
masonry) members, the yield curvature is essentially independent of reinforcement
content and axial load level, and is a function of yield strain and section depth alone.
This was discussed in relation to Fig.1.4. The form of the equation governing yield
curvature was given in Eq.(1.21). Based on the more extensive results presented in
Section 4.4, the following equations for yield curvature of some different section shapes
provide adequate approximations:
Circular concrete column: o,= (3.6a)
Rectangular concrete column: , (3.66)
Rectangular concrete wall: 9, (3.6c)
Symmetrical steel section: g, (3.6d)
Flanged concrete beam: g,= (3.66)
where & is the yield strain of the flexural reinforcement (={j/E;), and Dy te, Iy he and hy
are the section depths of the circular columa, rectangular column, rectangular wall, steel
section and flanged concrete beam sections respectively. Note that Eq.(3.6) gives the
curvature at the yield of the equivalent bi-linear approximation to the moment-curvature
curve, corresponding to point 3 on the force-displacement response in Fig. 1.6. As such it
is a useful reference value when using bi-linear force-displacement modelling.
For a SDOF vertical cantilever, such as a bridge pier, or a low rise cantilever wall, the
yield displacement can be satisfactorily approximated for design purposes by
9,(H + Lop) /3 co)
Y
For reinforced concrete and structural steel frames, as established in Section 4.4.6, the
yield drife can be developed from the yield curvature expressions of Eqs.(3.6) as
Reinforced concrete frame: 8, =0.5€,L, /h, (3.8)
.65€,L, Ih, 3.86)
Structural steel frame: 0, =
where Le is the beam span, and hg is the concrete or steel beam depth. It will be noted
that the yield drifts, and hence the yield displacements of reinforced concrete and7 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
structural steel frames with similar geometries differ only by 30%, and that concrete
frames are typically stiffer.
3.4.3 Equivalent Viscous Damping
‘The design procedure requires relationships between displacement ductility and
equivalent viscous damping, as shown in Fig. 3.1(¢). The damping is the sum of elastic
and hysteretic damping:
Seq = $01 + Sins: 69
where the hysteretic damping gis; depends on the hysteresis rule appropriate for the
structure being designed. Normally, for concrete structures, the elastic damping ratio is
taken as 0.05, related to critical damping. A lower value (typically 0.02) is often used for
steel structures.
Some discussion of both components of Eq.(3.9) is required.
(4) Hysteretic Damping: initial work on substitute-structure analysis (Le. analyses
using secant, rather than initial stiffness, and equivalent viscous damping to represent
hysteretic damping) by Jacobsenl'l, was based on equating the energy absorbed by
hysteretic steady-state cyclic response 10 a given displacement level 10 the equivalent
viscous damping of the substitute structure. This resulted in the following expression for
the equivalent viscous damping coefficient, Eyer
A,
Sips = TFA,
(3.10)
In Eq, (3.10), Ap is the area within one complete cycle of stabilized force-displacement
response, and Fy, and A are the maximum force and displacement achieved in the
stabilized loops. Note that the damping given by F.gs.(3.9) and (3.10) is expressed as the
fraction of critical damping, and is related to the secant stiffness Ke to maximum
response (see Fig.3.8). It is thus compatible with the assumptions of structural
characterization by stiffness and damping at peak response
Although this level of damping produced displacement predictions under seismic
excitation that were found to be in good agreement”!!! with time-history results for
systems with comparatively low energy absorption in the hysteretic response, such as the
modified Takeda rule, it was found to seriously overestimate the effective equivalent
viscous damping for systems with high energy absorption, such as elasto-plastic, or
bilinear rules. A reason for this can be found when considering the response of two
different systems with the same initial backbone curve (e.g, lines 1 and 2 in Fig.3.8) to an
earthquake record with a single strong velocity pulse, which might be considered an
extreme example of near-fault ground motion. Assume that one system has a bilinear8 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
analyses were separately averaged, and compared. In each case the equivalent viscous
damping was varied until the clastic results of the equivalent substitute structure matched
that of the real hysteretic model
F F
rk;
(a) Elasto-plastic (EPP) (b) Bi-linear, r= 0.2 (BI)
F
(c) Takeda “Thin” (TT) (a) Takeda “Fat” (TF)
ry
(c) Ramberg-Osgood (RO) (f) Flag Shaped (FS)
Fig.3.9 Hysteresis Rules Considered in Inelastic Time History Analysis'@
The hysteresis rules considered in the second study are described in Fig.3.9. The
clastic-perfectly plastic rule (Fig 3.9(@)) is characteristic of some isolation systems,
incorporating friction sliders. The biclinear elasto-plastic rule of Fig.3.9(b) had a second
slope stiffness ratio of r= 0.2, and is also appropriate for structures incorporating various
types of isolation systems, though the value of x can vary considerably. The two TakedaChapter 3. Dizect Displacement-Based Design: Fundamental Considerations 79
rules: Takeda Thin (Fig 3.9(c)) and Takeda fat (Fig. 3.9(4)), represent the response of
ductile reinforced concrete wall or column structures, and ductile reinforced concrete
frame structures respectively. Figure 3.9(¢) shows a bounded Ramberg Osgood rule
calibrated to represent ducule steel structures, and the flag-shaped rule of Fig.3.9(f)
represents unbonded post-tensioned structures with a small amount of additional
damping. Further information on these rules is provided in Section 4.9.2(g).
The two studies identified above ©.) initially were carried out without additional
clastic damping, for reasons that will become apparent in the following section. Figure
3.10 compares the resulting average relationships for an effective period of Te = 2.0
seconds for the four hysteresis rales common to both studies. It was found chat the
approaches resulted in remarkably similar relationships for equivalent viscous damping
for all hysteresis rules except clastic-perfectly plastic (EPP), where the discrepancy was
about 20%. It is felt that the difference for the EPP rule is a consequence of the use of
real records, with comparatively short durations of strong ground motion in [D1], and
artificial records, with longer strong ground motion durations in [G2]. It is known that
the EPP rule is sensitive to record duration, as the displacements tend to “crawl” in one
direction, particularly when P-A effects are incladed!"*.. Both studies showed the scatter
between results from different accelerograms to be greater for the EPP rule than for
other rules investigated. It is likely that the results from the [G2] study will be somewhat
conservative for shorter duration (i.e. lower magnitude earthquakes), but more realistic
for longer duration (higher magnitude) earthquakes. In the following discussion, che
average of the two studies has been used.
“The Dwairi and KowalskyiP" study represented the hysteretic component of response
in the form:
un
Siu = E (2) G11)
where the coefficient C depended on the hysteresis rule. This has an obvious relationship
to the theoretical area-based approach of Eig.(3.10) for the EPP rule, for which C= 2
Some period-dependency was found for effective periods T, < 1.0 seconds.
“The study by Grant et all, which considered a wider range of hysteretic rules, used a
more complex formulation of the relationship between ductility and equivalent viscous
damping, the hysteretic component of which is given by:
= Ly ! )
indi) aay] ee
Equation (3.12) includes the period-dependency of the response, in the coefficients ¢
and d. Table 3.1 lists the coefficients for the various hysteresis rules investigated.80
Priestley, Calvi and Kowalsky.
Displacement-Based Seismic Design of Structures
g 02 02
2 2
woa6 Boss
a &
£012 Son
a 3
2 0.08 200
© 0.04 5
3 a
=
BOP ooo o ToT
102 3 4 5 6 1 2 3 4 3 6
Displacement Ductility Displacement Ductility
(2) Elasto-Plastic (b) Thin Takeda
02 02
z z
S016 ‘0.16
5 &
Bon Gon
3 z
20.08 20.08
= 5
0.04 5.0.04
2 Ea
& a
° Too] °
1 2 3 4 35 6 1 2 3 4 35 6
Displacement Ductility
(©) Fat Takeda
Displacement Ductilicy
(2) Flag, B=0.38
Fig 3.10 Hysteretic Component of Equivalent Viscous Damping from Two
Independent Studies (D.K=Dwairi and Kowalsky!", GBP=Grant et all®2l)
Table 3.1 Equivalent Viscous Damping Coefficients for Hysteretic Damping
Component using Eq.(3.12)!@)
t Model a b c d
I EPP 0.224 | 0.336_| -0.002 [0.250
Bilinear, =0.2 (BN) 0.262 | 0.655 | 0.813 | 4.890
‘Takeda Thin (TT) 0.215 | 0.642 | 0.824 | 6.444
‘Takeda Fat (TF) 0.305 | 0.492 | 0.790 | 4.463
Flag, B=0.35 (FS) 0251 | 0.148 | 3.015 | 0511
Ramberg-Osgood (RO) 0.289 | 0.622_| 0.856 | 6.460Chapter 3. Direct Displacement-Based Design: Fundamental Considerations 81
‘The period-dependency part of Eq.(3.12) is plotted for the various hysteretic rules of
Fig.3.9 in Fig.3.11, where the hysteretic damping is related to the stable value found for
long periods, estimated at T= 4 sec. As with the other studi" the period dependency
was generally insignificant for periods greater than 1.0 seconds, with the EPP rule again
being the only exception. Since it is conservative to use low estimates of damping, and
since it will be unusual for regular structures such as frame and wall buildings, and bridges
to have effective periods less than 1.0 seconds, it will generally be adequate, and
conservative to ignore the period-dependency in design.
<4
Sry sersis:T
Enysteresissr/
Effective Period (seconds)
Fig.3.\1 Period Dependency of Hysteretie Component of Equivalent Viscous
Damping!@1
(b) Elastic Damping: Equation (3.9) includes an elastic component of equivalent
viscous damping, Elastic damping is used in inelastic time-history analysis (see Section
4.9.2(g)) to represent damping not captured by the hysteretic model adopted for the
analysis. This may be from the combination of a number of factors, of which the most
important is the typical simplifying assumption in the hysteretic model of perfectly linear
response in the elastic range (which therefore does not model damping associated with
the actual clastic non-linearity and hysteresis). Additional damping also results from
foundation compliance, foundation non-linearity and radiation damping, and additional
damping from interaction between structural and non-structural elements.
For single-degree-of-freedom (SDOF) systems, clastic damping is used in the dynamic
equation of equilibrium:
mi + ck + kx = mit, (3.13)82 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
where xis the response relative displacement, %, is the ground acceleration, m and k are
the mass and stiffiness, and the damping coefficient, cis given by
c= Imag = 2EVmk (3.14)
where @, = vk/m is the circular frequency, and €is the fraction of critical damping.
The damping coefficient, and hence the damping force depends on what value of
stiffness is adopted in Eq.(3.14). In most inelastic analyses, this has been taken as the
initial stiffness. This, however, results in large and spurious damping forces when the
response is inelastic, which, it is argued in Section 4.9.2@) is inappropriate, and that
tangent stiffness should be used as the basis for clastic damping calculations. With
tangent stiffness, che damping coefficient is proportionately changed every time the
stiffness changes, associated with yield, unloading or reloading, etc. This results in a
reduction in damping force as the structural stiffness softens following yield, and a
reduction in the energy absorbed by the elastic damping. Since the hysteretic rules are
invariably calibrated to model the full structural energy dissipation subsequent to onset of
yielding, this approach to characterization of the clastic damping is clearly more
appropriate than is initial-stiffness elastic damping. The significance to structural response
of using tangent-stiffness rather chan initial-stiffness damping is discussed in detail in
Section 4.9.2(9)
However, in DDBD, the initial elastic damping adopted in Fq.(3.14) is related to the
secant stiffness to maximum displacement, whereas it is normal in inelastic time-history
analysis to relate the elastic damping to the initial (clastic) stiffness, or more correctly, as
noted above, to a stiffness that varies as the structural stiffness degrades with inelastic
action (tangent stiffness). Since the response velocities of the “teal” and “substitute”
structures are expecred to be similar under seismic response, the damping forces of the
“real” and “substitute” structures, which are proportional to the product of the stiffness
and the velocity (Eq.(3.13)), will differ significantly, since the effective stiffness key of the
substitute structure is approximately equal to key =k; /H (for low post-yield stiffness).
Grant et all has determined the adjustment that would be needed to the value of the
elastic damping assumed in DDBD (based on either initial-stiffness or tangent-stifiness
proportional damping) to ensure compatibility between the “real” and “substitute”
structures. Without such an adjustment, the verification of DDBD by inelastic time-
history analysis would be based on incompatible assumptions of elastic damping,
‘The adjustments depend on whether initial-stiffness damping (conventional practice),
or tangent-stiffness damping (correct procedure, we believe) is adopted for time-history
analysis. If initial-stiffress damping is chosen, the elastic damping coefficient used in
DDBD must be larger than the specified initial-stiffiness damping coefficient; if tangent-
stiffness is chosen, it must be less than the specified tangent-stiffness coefficient.
This is explained further in Fig.3.12, which examines the stabilized tesponse (.e.,
ignoring the initial transient response) of a bilinear hysteretic model to steady-stateChapter 3. Direct Displacement-Based Design: Fundamental Considerations 83
harmonic excitation. The ¢ = constant line in Fig.3.12(b) represents the initial-stiffiness
damping assumption, and the dashed line the damping associated with the substitute
structure, if both aze assigned the same numeric value for damping, The energy dissipated
by clastic damping in the substitute structure is less than in the “real” structure (if we
accept initial-stiffness damping as correct), because key 1 sec (which will encompass
most designs, see Fig.3.11), and (b) an elastic damping ratio of 0.05 may be assumed. The
consequent ductility/damping relationships for the hysteretic rules considered are plotted
in Fig.3.14, for both tangent-stiffness and initial-stiffness elastic damping.
If we also acknowledge that the simplified Eq,(3.11) provides almost identical results
to the more complete expression of E.q.(3.12), if the period-dependency of Eq-(3.12) is
ignored, itis possible to include the ductility dependency of the elastic damping inside the
basic form of the equivalent viscous damping equations. The coefficients Cin q.(3.11)
are adjusted so that final value is correct with &, taken as 0.0586 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Steuctures
Equivatent Viscous Damping Ratio
6 T T T 1 6 T T T 1
° 2 4 6 8 ° 2 4 6 8
Displacement Ductility Pactor () Displacement Ductility Factor (i)
(@) Tangent Stiffness Elastic Damping (b) Initial Stiffness Elastic Damping
Fig.3.14 Design Equivalent Viscous Damping Ratios for 5% Elastic Damping
We list these equations below for tangent-stiffness elastic damping only, since this is
felt to be the correct structural simulation. Further, these equations cannot be altered to
apply for different levels of elastic damping by replacing the coefficient 0.05 by (say) 0.02,
since the coefficient C is valid only for & = 0.05. If different levels of &y are to be used,
the more compleze formulation of .q.(3.12) and Table 3.1 should be adopted. Where
appropriate, the corresponding structural type and material is identified in the following,
equations:
Conerete Wall Building, Bridges (TT): ,, = 0.05+0, “4 oa 7 ‘) (3.174)
Concrete Frame Building (IF): &, =0.05+0: ses # (3.17)
Steel Frame Building (RO) Sq = 0.05 sos (3.17)
Hybrid Prestressed Frame (FS, 8 =0.35): &, =0.05+0.18 a @G.17d)
Friction Slider (EPP): &,, = 0.05+0. oof 2) (3.17e)
Bilinear Isolation System (BI, r 054019 Hab (3.178)
Van
Note that the equations for the hybrid prestressed frame, and the bilinear isolation
system apply only for the parameters B=0.35 and r=0.2 respectively used in the analyses,
and will differ for other systems, Equations (3.{7d) and (3.176) should not be usedChapter 3, Direct Displacement-Based Design: Fundamental Considerations 87
without checking that the values of B or r are appropriate. This is addressed in more
detail in the relevant structural systems chapters.
Ideally, when hysteretic rules whose characteristics differ from those considered in the
above sections are used, an appropriate ductility/damping equation should be developed
based on inelastic time-history analyses (THA), in similar fashion to that described
above. It is recognized, however, that this will seldom be practical in a design
environment, though new equations are expected to be developed with on-going
research. Some reasonable estimates of the relationship can, however, be obtained by
comparing the relationships between the area-based viscous damping, given by Fq.(3.10)
which can readily be computed for any new system, provided the hysteretic response is
known, with the hysteretic component of the calculated viscous damping (Fig, 3.9, or
‘Table 3.1) for specific levels of the displacement ductility. This reladonship is plotted in
Fig. 3.15 for the six hysteresis rules considered above.
12
ITHA/Atea-Based EVD Damping Ratio
0 10 20 30 40 50
Area-Based Equivalent Viscous Damping(’)
Fig.3.15: Correction Factors to be Applied to Area-Based Equivalent Viscous
Damping Ratio (Eq.(3.10))
‘The vertical axis in Fig.3.15 is the ratio of the hysteretic component of the equivalent
viscous damping (EVD) found from time-history analysis, as reported above, to the arca-
based EVD from Fq,(3.10). This can thus be considered as a correction factor to be
applied to the area-based EVD. The three data points for each hysteresis rule correspond
to displacement ductilities of 2, 4, and 6. It will be noted that the trends are well
represented by the dashed lines for the three ductility levels, with no more scatter than
might be expected from the inherent scatter of the time-history results. Fig.3.15 can thus88 Priestley, Calvi and Kowalsky._Displacement-Based Seismic Design of Structures
be used to “correct” the EVD calculated from the area-based F.q.(3.10), for a known
displacement ductility level. Note that the elastic damping, in accordance with Fig.3.13 or
‘Table 3.2 then needs to be added to the hysteretic damping. This approach is suggested
as a suitable alternative to extensive time-history analyses for hysteretic rules that are not
represented by one of the equations (3.17).
(d) Example 3.2:
Fig. 3.16 Steady-State Harmonic Response of Hybrid Prestressed Structure with
r=0, and J=0.75 (Example 3.2)
As an example of this approach, consider the flag-shaped hysteresis loop of Fig.3.16, for
which r= 0, and J=0.75. ‘The area Ay of a complete cycle can be written as
A, =2x0.75F,(u-A,,
and hence, from Kq.3.10, the area-based EVD is
= 2X02, UDA, _ 075-1)
ow DAE, mm
For displacement ductilities of f= 2, 4 and 6, F.q.(3.18) yields Eiree = 0.119, 0.179 and
0.199 respectively. From Fig.3.15, the corresponding correction factors are 0.65, 0.76 and
0.88 respectively. Hence the hysteretic components of the EVD to be used in design are
(0.65x0.119=0,0774), (0.76x0,179=0.136), and (0.88x0.199=0.175) respectively. The
appropriate elastic damping from Fig.3.13 or Table 3.2 must then be added to these
values to obtain the total EVD to be used in design.
(3.18)Chapter 3. Direct Displacement-Based Design: Fundamental Considerations 89
(e) Generation of Inelastic Displacement Spectra
In Section 3.2 it was mentioned that inelastic displacement spectra sets could be
developed for different levels of displacement ductility as shown in the example of
Fig.3.3, If a relationship between initial-period elastic displacement and_ inelastic
displacement such as the equal-displacement approximation is assumed, the inelastic
spectra can be directly computed. Assuming that the skeleton force-displacement
response can be represented by a bi-linear approximation with a ratio of post-yield to
clastic stiffness equal to 4, the secant period T, to maximum displacement response is
related to the elastic period T; by the relationship
os
=7|—_“_
n (5) 49)
Since the inclastic displacement at T, must equal the elastic displacement at T; for the
equal displacement approximation to hold, and noting that elastic displacement response
is directly proportional to period, the modification factor Ry to be applied to the clastic
spectrum is
R -(ere) 620)
ue
Different relationships thus apply for different post-yield stiffnesses, but not to
different hysteretic energy absorption within the loop, provided that the equal-
displacement approximation is assumed to be valid. However, as is established in Section
49.2(@), the equal displacement approximation relies on the assumption that clastic
damping can be characterized by initial-stiffness proportional damping, which we have
demonstrated to be invalid.
Itis, however, possible to generate inelastic spectra sets directly from the data used «©
generate the damping-ductility reladonships of Fqs.(3.17). Substituting the reduction
factor for clastic damping values greater than 0.05 from Eq.(2.8) into the damping
ductility equations (Eqg.(3.17), spectral displacement reduction factors in the form
0.07 |
B=) Tact)
0.07+ f42)|
un
can be derived, Note that different relationships apply for different values of @ (the
coefficient dependent on whether “normal” or “velocity-pulse” conditions apply), for
different values of C and also, for different values of the elastic damping displacement
reduction relationship. As we have discussed in Section 2.2.3 there is still some
(3.21)90 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
uncertainty about the optimum form of this relationship. Consequently we believe it is
better to use, at this time, the more fundamental elastic displacement spectra set with
calibrated equations representing the damping/ductility relationship, than co combine
these in the form of Eq.(3.21).
For comparison, Fig3.17 plots the inelastic displacement spectra resulting from
Eqs.(3.20) and (3.21) for the elastic-perfectly plastic (EPP) hysteretic rule (ie. r=
[email protected]); € = 0.67 in Eq.(3.21). The differences between the two approaches are quite
considerable. It will be noted that for displacement ductilities of # 2 3 it would be
reasonable t0 use a constant reduction factor, simplifying design. We arrive at similar
conclusions for inelastic displacement spectra derived for other hysteretic rules.
os wl 0s wt
04
w=
na 03 =|
wt
geo 02
oa
0 °
Oo 1 2 3 0 4 5 o 12 3 4 5
Period (seconds) Period! (seconds)
(a) EPP (Equal Displacement) (b) EPP (Time-history Analysis)
Fig.3.17 Inelastic Displacement Spectra for Elastic-Perfectly Plastic Hysteresis
3.4.4 Design Base Shear Equation
Te will be clear that the approach described above can be simplified to a single design
equation, once the design displacement and damping have been determined. As noted
earlier, the displacement spectra are in many cases linear with effective period. The small
non-linearity at low periods is unlikely co be significant for displacement-based designs,
since it is the effective period at peak displacement response, approximately {2° times the
clastic period, that is of relevance. In Figs 3.1(d) and 2.5(b) the displacements are capped
at a petiod of 4 seconds, which in accordance with Eq.(2.3) might be considered
appropriate for an My = 6.9 earthquake. Let Ags be the displacement at the corner period
T. (e.g. Te =4 seconds in Fig. 3.1(d)) for the displacement spectrum corresponding t0 5%
damping. For 2 design displacement of Ay and design damping &, the effective period is,
from Fig. 3.1(d) and Eq.(2.8)
As (easy 6.22)
. A 0.07 )92 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
= 0.05+0.444(3.97-1)/3.97R = 0.155 (15.5%)
Maximum spectral displacement for 5% damping: ‘The corer period for peak
displacement response is T. = 4,0sec. Scaling to a PGA of 0.7 sec. from Fig. 3.1(d), which
applies for a PGA of 0.4g, the corresponding displacement is
Acs = 0.5x0.7/0.4 = 0.875m (34.4 in.)
Design Strength for Normal Ground Motion: Eauation (3.24) could be used directly
However, for clarity, the steps leading to Eq. (3.24) are taken sequentially. Applying che
damping correction factor of Fiq.(2.8) the corner-period response displacement for
15.5% damping is
{0.07
os
=0.875x| ———__|_ =0.553)
Beass (ots) ™
‘Thus, by proportion, the effective response period of the bridge is
T, = 4x0.35/0.553 = 2.53sec.
Note that Eq.(3.22) could have been used directly for the previous ewo steps
From Fq.(3.1), with the mass of (5000/g) tonnes, effective stiffness at peak response is
K, =4n?m,/T, =4n°5000/(9.805x2.537) = 3145 kN/m
Finally, from Eq. 3.2, the design base shear force is
Voge = K,A, =3145X0.35 = 1100 KN (247 kips)
Design Strength for Velocity Pulse Ground Motion: For this case, Eq. (3.24)
is used directly, with @ = 0.25:
‘ose
2 7 ozs
42°5000__ 0.875 (aes) = 1741 kN. (391 kips)
9.8054 0.35 \0.02+0.155,
‘This is 58% higher than for the normal ground condition case.
3.4.6 Design When the Displacement Capacity Exceeds the Spectral Demand
There will be occasions, with very tall or flexible structures, when the design
displacement capacity, calculated from Eqs.(3.4) or (3.5) exceeds the maximum possible
spectral displacement demand for the damping level calculated from Eq.(3.17). For
example, with reference to the response spectra set of Fig.3.1(d), it will be seen thac if the
design displacement Ay is calculated to be 0.35m, and the corresponding damping is 20%,
there is no possible intersection between the design displacement and the 20% damping
curve. At the corner period of 4 sec. the peak displacement in Fig.3.1(d) for 20%
damping is 0.282m. In such cases there are two possible conditions to be considered:Chapter 3. Direct Displacement-Based Design: Fundamental Considerations 93
(a) Yield Displacement Exceeds 5% Damping Value at the Corner Period: With
extremely flexible structures, or when the design seismic intensity is low, it is possible that
the yield displacement exceeds the 5% damping elastic response displacement (Ac,s) at the
comer period T, {in Fig.3.1(d) this is Ags = 0.5 m). In this case the calculated clastic
response period will be larger than 7,, the response displacement will be equai to Ags,
and the design base shear force is given by
Vea =Kabes (8.25)
Base
where Ky is the elastic stiffness. Note, however, that a unique design solution cannot be
found, since the stiffness, Ke depends on the elastic period, which depends, in turn, on
the strength. This is clarified in Example 3.4 below,
(b) Yield Displacement is Less than the 5% Damping Value at the Corner Period:
This case will be more common, Inelastic response will occur, but not at the level of
ductility corresponding to the displacement or drift capacity of the structure. Note that if
the yield displacement is less than Ags, this means that the elastic period is less than T.
As the structure softens, a final effective period of T2T, will be achieved, with a
displacement response level that is compatible with the damping implied by that
displacement. ‘The following trial and error solution method is recommended:
1. Calculate the displacement capacity Ay, and the corresponding damping, &.
Confirm that the two are incompatible with the displacement spectra set, as
above
2. Estimate the final displacement response Ay. ‘This will be somewhere between
Acgeand Aye
Calculate the displacement ductility demand corresponding to Ay: (= Ay/ Ay)
Calculate the damping & corresponding to the ductility demand
Calculate the displacement response A at J; corresponding to &
Use this value A as the new estimate for the final displacement Aye.
Cycle steps 3 to 6 until a stabie solution is found. ‘Typically this requires only one
or owo iterations.
Again there is no unique solution, as the effective stiffness could correspond to any
period T > T,. Any value of design base shear less than Vaase = 47?m-Ay//T,’ will satisfy
the design assumptions. A higher value will imply a response at an effective period less
HO ae
than Ts, and hence the response displacement will be incompatible with the effective
damping. In both cases discussed above the provided strength will nor affect the
displacement response. Minimum strength requirements for P-A effects (Section 3.6) or
gravity loads will govern the required strength,
3.4.7 Example 3.4: Base Shear for a Flexible Bridge Pier
Example 3.3 is redesigned for somewhat different conditions. First, (case (a)), the 10m
(32.8f) high bridge pier is designed for normal ground motion with a reduced peak94 Priestley, Calvi and Kowalsky, Displacement-Based Seismic Design of Structures
ground acceleration of 0.35g, and a spectral displacement spectra set proportional to
Fig.3.1(d). Second, (case (b)), the physical dimensions of the column cross section are
maintained, but the effective height is increased co 25m (82f).
Case (a): Design Displacement: Limi state conditions have not changed from
Example 3.3, so the design displacement capacity is still 0.35m (13.8 in.), and the
corresponding displacement ductility capacity is f =3.97. The corresponding equivalent
viscous damping at the design displacement capacity is again 15.5%.
Maximum spectral displacement for 5% damping: With a corner period of 4.0
sec., and 0.35g PGA, the corner-period elastic displacement is scaled from Fig.3.1(d)
which applies for a PGA of 0.4g to give Acs = 0.5x0.35/0.4 = 0.438 m (17.2 in.). This
exceeds the yield displacement of 0.0881, so the pier will respond inelastically.
Maximum spectral displacement for 15.5% damping: By proportion from
Example 3.3, the corner displacement for 15.5% damping is:
Agyes = 0.533x0.38 /0.7 = 0.277m
‘This is less than the displacement capacity of 0.35m, and hence the response
displacement will be less than the displacement capacity. We use the itetative approach
outlined in Section 3.4.7(b):
2. The final displacement will be somewhere between the corner displacement, 0.277 m
and the displacement capacity, 0,35 m. We make an initial estimate of Ay = 0.30 m.
3. With the yield displacement at A, = 0.0881 m (see Example 3.3) the displacement
ductility is = 0.30/0.0881 = 3.41
4, From Fiq.(3.17a) the corresponding equivalent viscous damping ratio is
GA4l-
= 0.05 + 0.444
$=0.05+ Sar
0.150 (15%)
5. The corresponding displacement at the corner period 7; is again found from Eq.2.8):
(0.07
Ay, =0.438-| —
0.02+ 0.15
Use this as the new estimate of Aa.
6. Cycling once more through the sequence yields: f= 3.19, & = 0.147, and Axg = 0.284
m (11.2in.). The result has stabilized.
‘The reference effective stiffness is thus found, using T= T.= 4.0 sec. in Eq.3.1) as:
K, = 427 -(5000/9.805)/4? = 1258 kN/m ,
os
] = 0.28)
and the maximum design base shear force, from Eq.(3.2) as
Vesose = 1258X0.284 = 357.3KN. (80.3 kips)
Gravity load or P-A requirements will govern the choice of the actual design strength.Chapte: 3. Direct Displacement-Based Design: Fundamental Considerations 95
Case (b): Design Displacement: The yield curvature is unchanged, and the yield
displacement is calculated, ftom Eq.(3.7), as
Ay = BHP /3 = 0.00264x252/3 = 0.55 m (21.7 in).
Note that this is independent of the final strength. This displacement exceeds the elastic
displacement at the corner period, Ays = 0.438 m. Hence there is no point in further
calculating the displacement capacity, as the pier will respond elastically to the design level
of intensity, with a response displacement of 0.438 m.
Now, since the structure responds elastically, with a known displacement, the allocated
strength is in fact arbitrary, as noted above. For example, if we allocated a yield strength
of 500 kN, the stiffness would be Ky = 500/0.55 =909 kN/m. ‘The calculated elastic
period would be:
T,, = 2J(5000(9.805x909)) = 4.71 seconds.
‘The response displacement would be 0.438 m, and the maximum response force would
be
Veave = 909X0.438 = 398 KN (89.5 kips).
However, if we arbitrarily allocated a yield strength of 350KN, the stiffness would be
Kq =350/0.55 = 636 KN/m, and the elastic period would be 5.63 seconds. ‘The response
displacement, (see Fig.3.1(d)), would still be 0.438 m, and the maximum response force
would be 278 KN (62.5 kips). P-A moments for this case would be 36% of the base
moment from the horizontal inertia force, and, in accordance with Section 3.6, would
need to be carefully considered,
Nore that if the strength was arbitrarily set higher than 692 kN, (say 800 kN), then the
elastic period would be found to be less than 4 sec. (in this case, Ka =1455 KN/m, and
Tey = 3.72 sec, and the structure would respond with a displacement less than 0.438 m, (in
this case 0.407 m (16.0 in)), in accordance with the clastic 5% displacement spectrum for
the calculated period.
In fact, the example is probably artificial. The base moment would be very high, in
cither case, and redesign with a larger column diameter (and hence smaller yield
displacement) would be advisable. The pier would also be excessively flexible for gravity
loads.
3.5 MULTI-DEGREE-OF-FREEDOM STRUCTURES
For multi-degree-of-freedom (MDOF) structures the initial part of the design process
requires the determination of the characteristics of the equivalent SDOF substitute
siructurd\, The required characteristics are the equivalent mass, the design displacement,
and the effective damping, When these have been determined, the design base shear for
she substitute structure can be determined, The base shear is then distributed between the
mass elements of the real structure as inertia forces, and the structure analyzed under
these forces to determine the design moments at locations of potential plastic hinges.96 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
3.5.1 Design Displacement
‘The characteristic design displacement of the substitute structure depends on the limit
state displacement or drift of the most critical member of the real structure, and an
assumed displacement shape for the structure. This displacement shape is that which
corresponds to the inelastic first-mode at the design level of seismic excitation. Thus the
changes to the elastic first-mode shape resulting from local changes to member stiffness
caused by inelastic action in plastic hinges are taken into account at the beginning of the
design. Representing the displacement by the inelastic rather than the elastic first-mode
shape is consistent with characterizing the structure by its secant stiffness to maximum
response. In fact, the inelastic and elastic first-mode shapes are often very similar.
‘The design displacement (generalized displacement coordinate) is thus given by
Awe Thay T oma) 3.26)
where m; and A; are the masses and displacements of the 1 significant mass locations
respectively. For multi-storey buildings, these will normally be at the # floors of the
building. For bridges, the mass locations will normally be at the centre of the mass of the
superstructure above each columa, but the superstructure mass may be discretized to
more than one mass per span to improve validity of simulation (see Section 4.9.2(¢) ii).
With tall columns, such as may occur in deep valley crossings, the column may also be
discretized into multiple elements and masses.
Where strain limits govern, the design displacement of the critical member can be
determined using the approach outlined in Section 3.4.1. Similar conclusions apply when
code drift limits apply. For example, the design displacement for frame buildings will
normally be governed by drift limits in the lower storeys of the building. For a bridge, the
design displacement will normally be governed by the plastic rotation capacity of the
shortest column. With a knowledge of the displacement of the critical element and the
design displacement shape (discussed further in the following section), the displacements
of the individual masses are given by
A.
A, =6, {#] (3.27)
(Oc.
where gis the inelastic mode shape, and Ae is the design displacement at the critical
mass, 6 and 6s the value of the mode shape at mass &
Note that the influence of higher modes on the displacement and drift envelopes is
generally small, and is not considered at this stage in the design. However, for buildings
higher than (say) ten storeys, dynamic amplification of drift may be important, and the98. Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Syn Gin
Height Ratio (H\/H,)
be = 265 /,
(@) Curvature 0 02 04 06 08 1
(6) Displacement Ratio
Fig.3.18 Yield and Design Displacements for Cantilever Walls
If the roof drift from Hq (3.30) is less than the code drift limit ®., chen the design
displacement profile is given by
H, 2e, )
}}1-—+ |+| ¢, -— |L,H, 32:
(1-2) +0. a ce)
/
A, =A, +4,, =
If the code drift limit governs the roef drift, the design displacement profile is given by
H,
A,=A,+(0-6,,)H, oH (! # (2, 32)
Although Eq, (3.32) can be manipulated to provide a generalized displacement shape &
to be compatible with Eq. (3.27), there is little value in so doing, since the full
displacement profile must first be found. Further information on displacement profiles
for cantilever walls is provided in Chapter 6.
(©) Multi-Span Bridges: With bridges it is less easy to initially determine a design
displacement profile, particularly for transverse seismic response. Figure 3.19 illustrates
two possible bridge configurations out of a limitless potential range. The example of Fig.
3.19(a) has piers of uniform height, while those in Fig. 3.19(b) vary in height. The
transverse displacement profiles will depend strongly on the relative column stiffnesses,
and more significantly, on the degree of lateral displacement restraint provided at the
abutment, and the superstructure lateral stiffness. For each bridge type, three possible
transverse displacement profiles are shown, corresponding to an abutment fully
restrained against transverse displacement, a completely unrestrained abutment, and one
where the abutment is restrained, but has significant transverse flexibility.Chapter 3. Direct Displacement-Based Design: Fundamental Considerations 99
x Lay por
lt
Structure Structure
12
apes sl
208
= a6
boa
estained 20.2 fresiaines
TTT ° TT TT
0 02 04 06 08 1 0 02 04 06 08 1
Dimensionless Distance Dimensionless Distance
() Uniform Height Piers (b) Irregular Height Piers
Fig.3.19 Design Transverse Displacement Profiles for Bridges
For the case of Fig, 3.19(a), the critical pier will be the central one, and with the
appropriate displacement profile chosen, Eq, (3.27 and 3.26) can be applied directly. For
the irregular bridge of Fig. 3.19(b) the critical pier may not be immediately apparent, and
some iteration may be required. Iteration may also be required for the case of finite
translational flexibility of the abutments for both the regular and irregular bridges +o
determine the relative displacements of abutment and the critical pier. Generally a
parabolic displacement shape between abutments and piers can be assumed for initial
design. Further information on displacement profiles for bridges is given in Chapter 10.
3.5.3 Effective Mass
From consideration of the mass participating in the first inelastic mode of vibration,
the effective system mass for the substitute structure is
m, =D (mA,)iA, (333)
ai
where Ay is the design displacement given by Eq.(3.26). Typically, the effective mass will
range from about 70% of the total mass for multi-storey cantilever walls to more than
85% for frame buildings of more than 20 storeys. For simple multi-span bridges the
effective mass will often exceed 95% of the total mass. The remainder of the mass
participates in the higher modes of vibration, Although modal combination rules, such as100 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
the square-root-sunvof squares (RSS) ot complete quadratic combination (CQC)rules: may
indicate a significant increase in the elastic base shear force over that from the first
inelastic mode, there is much less influence on the design base overturning moment. The
effects of higher modes are inadequately represented by elastic analyses, as will be shown
in the chapters devoted to specific structural forms, and are better accommodated in the
capacity design phase, rather than the preliminary phase of design
3.5.4. Equivalent Viscous Damping
(a) System Damping: The effective damping depends on the structural system and
displacement ductility demand, as illustrated in Fig.3.1(0) and Eqs.3.17. This requires
determination of the displacement ductility demand of the substitute structure, This poses
few problems, since the design displacement Ay has already been determined, from
Eq,(3.26). The effective yield displacement A, needs to be interpolated from the profile
of displacements at yield (¢.g, Eq.(3.31) for cantilever walls, or Eq.(3.8) for frames). For
frames it is adequate to assume that the yield drift is constant with height (i.e. the yield
displacement profile is linear with height), and hence the yield displacement is
A, =6,.H, (G34)
where 6, is given by Eq.3.8). For walls, the yield displacement is found from Eq.(3.31)
with H; = He. In both cases this requires knowledge of the effective height of the
substitute structure, which may be taken as:
H= DY (mA) (mA) (3.35)
ia 7
The design ductility factor, for use in Eq.(3.17) is then
u=A,/A, 3.36)
in the usual fashion.
Note that provided reasonable ductility is implied by the design displacement Ag,
Fig.3.1(@) and [email protected]) indicate that the damping is not strongly dependent on the
ductility, and average values may be adopted. This is also implied in Fig.3.17(b). Note
also, that concrete and masonry structures are much more flexible than normally assumed
by designers, and hence code drift limits, rather than displacement ductility capacity tends
to govern design (sce Section 5.3.1, eg,). As a consequence, the design ductility, and the
effective damping are known at the start of the design process, and ao iteration is needed
in determining the design base shear force.
When the lateral resistance of a building in a given direction is provided by 2 number
of walls of different length, the ductility demand of each wall will differ, since the yield
displacements of the walls will be inversely proportional to the wall lengths (see
Eq.(3.6c)), while the maximum displacements at design-level response will be essentiallyChapter 3. Direct Displacement-Based Design: Fundamental Considerations 101
equal, subject only 1 small variations resulting from torsional response and floor
diaphragm flexibility. ‘This was discussed in Section 1.3.6, with reference 10 Fig, 1.13
Hence the system damping will need to consider the different effective damping in each
wall
In the general case, where different structural elements with different strengths and
damping factors contribute to the seismic resistance, the global damping may be found by
rhe weighted average based on the energy dissipated by the different structural elements.
That is,
é La, é, von, 637
where Vj, Aj and & are the design strength at the design displacement, displacement at
height of centre of seismic force, and damping, respectively, of the # structural element.
Alternatively, the energy dissipated may be related to the moment and rotation of
different plastic hinges (VjA; =MjQ). This form may be more appropriate for frame
structures.
With multiple in-plane walls, the displacements of the different walls will all be the
same, and hence Eq,(3.37) can be simplified to
é. xe! Ly, 638)
where Vj and are the base shear force and damping of the m walls in a given direction
Some modification of Eg.(3.38) may be required when torsional response of a building
containing more than one plane of walls in a given direction is considered. In this case,
F-q.(3.37) applies. However, the error involved in using Eq.(3.38) is small, even when
torsional response is expected.
A tational decision will be to apportion the total base shear force requirement
derween the walls in proportion to the square of the length. This will result in essentially
constant reinforcement ratios between the walls. With wall strength proportional to
length squared, Fq.(3.38) may be rewritten 2s:
b= Zlte P NE 3.39)
(8) Influence of Foundation Flexibility on Effective Damping: Although the
influence of foundation flexibility on seismic design can be incorporated into force-based
design, albeit with some difficulty, it is rarely considered. Foundation flexibility will
increase the initial elastic period, and reduce the ductility capacity corresponding to the
strain or drift limic statesl*!, It is comparatively straightforward, however, to incorporate
the influence of elastic foundation compliance into Direct Displacement Based Design. li
the limit state being considered is strain-limited, then the design displacement will be
increased by the elastic displacement corresponding to foundation compliance (this102 Priestley, Calvi and Kowalsky.
Jacement-Based Seismic Design of Structures
requires a knowledge of the design base moment and shear force, and hence some
iteration may be required). If, however, the limit state is defined by code drift limits, there
will be no change in the design displacement, thus implying reduced permissible
structural deformation,
Vine
(a) Foundation (b) Structure (c) Foundation + Structure
Fig.3.20 Damping Contributions of Foundation and Structure
The second influence relates to the effective damping. Both foundation and structure
will contribute to the damping. Consider the force-displacement hysteresis loops of
Fig.3.20, where foundation (A) and structure (A) components of the peak response
displacement Ay = 4s + Ay have been separated for a cantilever wall building. Assuming
sinusoidal displacement response, the area-based equivalent viscous damping for the
fonndation and for the structure can be separately expressed as
A
i -_ 4
Foundation: ere 3.408)
Sree 2aW,,.A, (3.402)
A
Structure: buna = 5 (3.40b)
OV ah,
where Ay and A, are hysteretic areas within the loops (Le. energy absorbed per cycle) for
foundation and structure respectively. As shown in Fig,3.20, the hysteretic area of the
combined structure/foundation system will be the sum of the two components, and
hence the system equivalent viscous damping will be
A+4, SA +E,
System: (3.406)
~ AV gal, +A, A, +A,104 Priestey, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Table 3.3 Calculations for Example 3.5
Height Hi (m) a aa HA, |
3.2. 0.0482 0.00232 0.1542
6.4 0.1041 0.01084 0.6662
Cae 0.1651 0.02726, 1.5850
Rs 0.2287 0.05230 2.9270
Sum = 0.5461 0.09272 5.332
From Eq.(3.26), noting that the masses are all equal, the design displacement of the
SDOF substitute structure is
A, = D(mai vee (m,A,) = 0.09272/0.5461 = 0.1698m (6.68 in.)
a
Yield displacements:
by E435):
he effective height of the SDOF substitute structure is given
H, xt m.A,H, Dx 'm,A, )=5.332/0.5461 =9.765m (=0.763.H,)
mT
Yield displacements of the 4 m and 2 m walls are given by Eq.(3.31). For the 4 m wall:
) f
é, wr) 000 x9.765° 1— ae) = 0.0355m (1.40 in.)
4
3H, 3x128
‘The yield displacement of the 2m wall will be twice this, ic. 0.0710m (2.80in).
Wall displacement ductility factors: From §q.(3.36), the displacement ductility
factors for the walls are:
4-m wall: = 0.1698/0.0355 = 4.78
2m wall: {£= 0.1698/0.0710 = 2.39
Wall damping factors: From Eq.(3.17a), the individual wall damping factors are:
4m wall: , = 0.05+0.444(4.78 - 1)/(4.78m) = 0.162
2 m wall: & = 0.05+0.444(2.39 - 1)/(2.39m) = 0.132
Structure equivalent viscous damping: Finally, assuming a distribution of base
shear force in proportion to (wall length)?, as discussed in section 3.5.4, then, from
Eq.(3.39), the effective damping, to be used in design is:
((222)x0.132 +42x0.162)/(2x2? + 42) = 0.152 (15.2%)
3.5.6 Distribution of Design Base Shear Force
The principles outlined i the previous sections enable the design base shear to be
established for a MDOF system. This base shear force must be distributed as designChapter 3. Direct Displacement-Based Design: Fundamental Considerations 105
forces to the various discretized masses of the structure, in order that the design
moments for potential plastic hinges can be established. Assuming essentially sinusoidal
response at peak response, the base shear should be distributed in proportion to mass
and displacement a¢ the discretized mass locations. Thus the design force at mass /is:
F,=VyaelmA,) SA.) oan
Similarity with force-based design for multi-storey buildings will immediately be
apparent. The difference is that the design inelastic displacement profile, rather than a
height-proportional displacement is used, and the form of Eq,(3.41) is generalized co all
structures (including, eg. bridges and wharves), not just buildings. It will be shown in
Chapter 5 that minor modification of Hq,(3.41) is advisable for taller frame buildings to
avoid excessive drift in upper storeys.
‘The distribution of the design base shear force between different parallel lateral force-
resisting elements (walls, and frames) is, to some extent, a design choice, as will be
discussed in more detail in individual chapters relating to different structural systems.
3.8.7 Analysis of Structure under Design Forces
Analysis of the structure under the lateral force vector represented by Eq.3.41) to
determine the design moments at potential plastic hinge locations is analytically
straightforward, but nevertheless needs some conceptual consideration, In order to be
compatible with the substitute structure concept that forms the basis of DDBD, member
stiffness should be representative of effective secant stiffnesses at design displacement
response
For cantilever wall buildings, this can be simplified to a distribution of the vertical
force vector between walls in proportion to J,’, as suggested above, with the walls then
analysed separately. The designer should not, however, feel unduly constrained by this,
suggested strength distribution, as there will be cases where the adoption of other
distributions is more rational. This is discussed further in Chapter 6.
For reinforced concrete frame and dual (wall/frame) system buildings, more care is
needed. With weak-beam/strong-column frame designs, beam members will be subjected
10 inelastic actions, and the appropriate beam stiffness will be:
(ED) pean = EL, | My (3.42)
where Edler is the cracked-section stiffness, found in accordance with the methods
developed in Chapter 4, and ff is the expected beam displacement ductility demand
\nalyses have shown that the member forces are not particularly sensitive to the level of
stiffness assumed, and thus it is acceptable to assume that flp=Hh, the frame design
ductility.
Since the columns will be protected against inelastic action by capacity design
procedures, thelr stiffness should be taken as Exley with no reduction for ductility. NoteChapter 3. Direct Displacement-Based Design: Fundamental Considerations 1
Note chat this is 36% higher than required for the damage-control limit state of
Design Example 3.6, and 256% higher than required for Case 1, when the drift limit was
effectively 0.01. This illustrates the very non-linear nature of the strength/displacement
requirement noted subsequently in Section 3.10.1. Since the 2m walls will only develop
about 50% of their nominal flexural strength at the serviceability design drift limit, the
total nominal base shear capacity will need to be larger than 677KN (in fact, about
810KN). It should also be noted, however, that the low drift limit of 0.005 has been
imposed because of the requirement to limit damage 10 masonry infill panels. In the
initial stages of response, the infill will provide substantial additional lateral resistance to
the building, which should be considered when determining the required strength of the
cantilever walls. Since the infill walls are expected to be severely damaged at the damage
control limit state, their contribution to lateral resistance at this level of response would
be discounted. It is thus quite possible that the structure would also satisfy the required
strength for this rather stringent serviceability limit state if designed to the damage-
control limit state. Design considering masonsy infill is discussed in Section 5.12,
3.6 P-A EFFECTS
3.6.1. Current Design Approaches
As structures displace laterally, as suggested, for example, in the single-degree-of-
freedom approximation of Fig.3.23(a), gravity loads induce overturning moments in
addition to those resulting from lateral inertia forces, Using the nomenclature of Fig.3.23,
the base moment Mis
M=FH+PA (3.43)
P
Ay? Strength Enhanced for PA
>
F F|
2
2
5
H z
| PA -FH| Ay Displacement Ay
(a) Structure (b) Moments (c) Force-Displacement Response
Fig.3.23 P-A Effects on Design Moments and Response12 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structutes
If the base moment capacity Mp is developed in inelastic response, then the lateral
inertia force that can be resisted reduces as the displacement increases, according to the
relationship
Mp-PA
H
F= 3.44)
This effect is illustrated in Fig.3.23(c), where it is apparent that the P-A effect not only
reduces the lateral force, but also modifies the entire lateral force-displacement
characteristic. ‘The effective initial stiffness is reduced, and the post-yield stiffness may
become negative.
The significance of P-A effects is recognized in most seismic design codes, and is
gypically quantified by some form of “stability index”, 8, which compares the magnitude
of the P-A effect at either nominal yield, or at expected maximum displacement, to the
design base moment capacity of the structure. Since the P-A effect is of maximum
significance at the design level of seismic response, we relate the stability index to
conditions at maximum response, as recommended in [PI]:
PA
nx 5)
% M, (3.45)
In conventional force-based design, one of two different approaches is typically
adopted to account for P-A effects. One approach is to increase the expected design
displacement to A* mux
Mad,
a, 1-8
‘The alternate approach is to increase the strength in an attempt to avoid an increase in
the expected design displacement. Paulay and Priestleyi", discussing the design of
reinforced concrete frame buildings, recommend that when the stability index is less than
)=0.085, P-A effects may be ignored. For higher values of the stability index, an equal-
energy approach is adopted to determine the required strength increase, as suggested in
Fig, 3.23(©). This implies that the required nominal strength increase, ignoring P-A effects
is somewhat greater than 50% of the calculated P-A effect.
3.46)
3.6.2. Theoretical Considerations
Inelastic time-history analyses(®29I indicate that the significance of P-A effects
depends on the shape of the hysteretic response. With the adoption of an elasto-plastic
characteristic it can be shown™' that if the earthquake record is long enough, instability13
Chapter 3. Direct Displacement-Based Design: Fundamental Consider:
i.e, an increase in displacements until PA=Mp) will eventually occur, when P-A effects
are included in the dynamic analysis. This can be explained with reference to Fig. 3.24(a),
After an initial inelastic pulse resulting in a maximum displacement corresponding to
point A, the structure unloads down a line of stiffness equal to the initial elastic stiffness
co point B. In further response cycles it is more probable that the strength envelope will
be reached at point A rather than point C on the opposite yield boundary, since a higher
ievel of elastic response is needed to attain point C. Consequently, once the first inelastic
pulse occurs, it creates a tendency for continued displacement in the same sense, and
response continues incrementally to D and E, and if the earthquake record is long
enough, failure eventually occurs,
Although elasto-plastic hysteretic characteristics may be a reasonable approximation to
steel structure response, concrete structures are better represented by the modified
Takeda hysteretic rulel|, illustrated in Fig. 3.24(b), which has a positive post-yield
stiffness, an unloading stiffness that is significantly less than the initial loading stiffness,
and subsequent re-loading stiffness greatly reduced from the initial stiffness. ‘The positive
post-yield stiffness compensates, to some extent for the strength loss associated with P-A.
moments. Furthermore, unloading from the same displacement (point A) as for the
elasto-plastic case results in much lower residual displacement at point B, because of the
reduced unloading stiffness. Subsequent elastic cycles result in gradual reduction of the
residual displacement (shake-down, lines B-F-G in Fig. 3.24(b)) due co the reduced
stiffness in the reverse direction, and no preferential direction for cumulative
displacement develops.
(a) Elasto-plastic Hysteresis (b) Takeda Degrading Stiffness Hysteresis
Fig 3.24 Influence of Hysteresis Rule on P-A Response114 Priestley, Calvi and Kowalsky, Displacement-Based Seismic Design of Structures
Analyses have shown! that provided the second slope stiffness K", of the stabilized
loop shape, including P-A effects is positive, as in Fig. 3.24(b), structural response is
stable, with only minor increase in displacement compared with response where P-4
effects are ignored. it has been shown!" that for Takeda response, a positive
stabilized second slope stiffness of at least 5% of the initial elastic stiffness is assured
provided the stability index satisfies:
4, 50.3 G47)
where the stability index is defined by Eq.(3.45)..
3.6.3. Design Recommendations for Direct Displacement-Based Design
It will be recognized that there are significant difficulties in rationally considering P-A
effects in force based design. As was noted in Section 1.3.7, estimation of maximum
expected displacement from different codified force-based designs is subject to wide
variability, Hence large ertors in calculated P-A moments can be expected, depending on
which design code is used. Further, most force-hased codes seriously underestimate the
clastic and inelastic displacements, and hence underestimate the severity of P-A effects.
The treatment of P-A effects in DDBD is comparatively straightforward, and is
illustrated in Fig, 3.23(€), Unlike conditions for force-based design, the design
displacement is known at the start of the design process, and hence the P-A moment is
also known before the required strength is determined. DDBD is based on the effective
stiffness at maximum design displacement, When P-A moments ate significant, it is the
stiffness corresponding to the degraded strength and the design displacement (see Ke in
Fig.3.23(¢)) that must match the required stiffness. Hence, Eq.(3.2) defines the required
residual strength. The initial strength, corresponding to zero displacement, is thus given
by
PA,
F=KA,+C- 48)
and hence the requited base-moment capacity is
M,=KA,H+C-PA, (3.49)
Note that it is more consistent to define the P-A effect in terms of the base moment,
than the equivalent lateral force. In Eq,(3.49), for consistency with the design philosophy
of DDBD, we should take C=1. However, examination of the hysteretic loops indicates
that more energy will be absorbed, for a given final design displacement and degraded
strength, than for a design when P-A design is not required, particularly for conceete-like
response. It is also apparent from time-history analyses that for small values of the
stability index, displacements ate only slightly increased when P-A moments are ignored,Chapter 3. Direct Displacement-Based Design: Fundamental Considerations 5
as noted above, It is also clear that steel structures are likely to be more critically affected
than will concrete structures. Consideration of these points leads to the following design
recommendations, which have recently been confirmed by time-history analyses!”
(4) Steel structures: When the structural stability index defined by Eq. 3.45 exceeds
0.05, the design base moment capacity should be amplified for P-A considerations as
indicated in F.9.3.49), taking C=1. This is represented by line 1 in Fig 3.25. Note that
this implies greater strength enhancement than indicated by the upper line in Fig.3.23(c)
For lesser values of the stability index, P-A effects may be ignored
(b) Concrete structures: When the structural stability index defined by F9.(3.45)
exceeds 0.10, the design base moment capacity should be amplified for P-A effects as
indicated in Bq.(3.49), taking C=0.5. This is represented by line 2 in Fig.3.25. This
” in Fig. 3.200)
corresponds to the upper line, marked “strength enhancement for P-
For lesser values of the stability index, P-A effects may be ignored
For both steel and concrete structures, it is recommended that the Stability Index,
given by Eq.(3.45) should not exceed 0.33.
Force
(Ke + P/H)Ag
(K. + 0.5P/H)Ay
Keds Steet
Concrete
Displacement Aa
Fig. 3.25 Required P-A Strength enhancement in Displacement-Based Design
3.7 COMBINATION OF SEISMIC AND GRAVITY ACTIONS
3.7.1 A Discussion of Current Force-Based Design Approaches
Force-based seismic design codes normally require that actions (moments and shears)
resulting from seismic design forces (reduced from the elastic level by specified force-116 Priestley, Calvi and Kowalsky. Displacement-Based Seissnic Design of Simmecures
reduction factors) be directly added to gravity moments and shears to determine the
required design strength. Since this can result in very unbalanced moment demands at
different critical locations of a structure, limited moment redistribution is advocated in
some design texts (e.g;[PI]), and permitted in some design codes to improve structural
efficiency. There are a number of illogical aspects related to the current philosophy for
combination of gravity and seismic actions, and these will be examined before making
design recommendations.
Consider the bridge bent subjected to gravity loads (G) from owo bridge girders, and
seismic lateral forces (B), illustrated in Fig, 3.26. The columns are circular, and under
seismic response, che normal design philosophy is adopted, that plastic hinges should
form only in the columns and not in the cap beam. Gravity load moments will often be
determined from an analysis assuming gross (un-cracked) section stiffness. The results
from such an analysis are shown by the solid line in Fig, 3.26(b), with key relative
magnitudes included in parentheses.
Seismic moments may well be calculated using different stiffness assumptions —
typically the column stiffnesses will be reduced to take some allowance for the effects of
cracking, at moments corresponding to the yield moment (see Section 4.4). It should be
noted, however, that combining results from different analyses using different stiffness
values violates compatibility requirements. The results of such an analysis, shown in Fig,
3.26(c), corresponding to the elastic moments (Mg) from a response-spectrum analysis,
reduced by the design force-reduction factor, R, (1.e., Meseduces =Me/R) are significantly
larger than the gravity moments, as would be expected from a region of moderate
seismicity. Relative moment magnitudes are again shown in parentheses in Fig.3.26(C).
In the following discussion, it is assumed that the “equal displacements”
approximation is reasonably valid (but see Section 4.9,2(g) for a discussion of this
assumption), and hence the displacement ductility factor ly = R.
The gravity and reduced seismic moments are combined in Fig. 3.26(d). It is seen that
very different moment demands are created in the two columns, and thar the critical
locations are the top (A) and bottom (B) of the right columa, with moments of 15 and 13
units respectively.
In the following, the influence of seismic axial load, which will affect the moment
capacity of the nwo columns by different amounts will be ignored. The first consideration
to be made is that designing for these combined magnitudes will result in an inefficient
structure. Under reversal of the direction of the seismic force JB, the left hand column will
become critical, and hence both columns will have to be designed for the same flexural
strength. We assume that the columns are cizcular, and that the moment capacities of the
potential plastic hinges are equal in both directions of loading. Under seismic response,
the pattern of moments shown in Fig. 3.26(d) by the solid line may be a reasonable
approximation at first yield of the bent, but as the structure continues to deform to
maximum displacement response, the left hand column wil! quickly develop plastic hinges
at top and bottom, at che design strength of 15 units, assuming the reinforcement is the
same at top and bottom of the columns. The pattern of moments throughout the118 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
structure when a full plastic hinge mechanism has formed is shown in Fig, 3.26(a) by the
dashed line.
At first vield, for the direction of seismic force indicated in Fig.3.26(d), the right-hand
column carries a shear force of (15+13)/H, where His the column height, while the left-
hand column carries a shear of only (3+7)/#L The total seismic force is thus 38/1. Note
that this is also the value obtained directly from the design seismic moments of
Fig.3.26(c), without consideration of the gravity loads. However, when the full
mechanism forms, each column carries (15+15)/Ff, for a total of 60/H. The actual
lateral strength developed will thus be 60/38 = 1.58 times the design lateral force used to
determine the design seismic moments. If section strength reduction factors are adopted
and conservative estimates are made of material strengths, then the probable lateral
strength may well exceed nwo times the design lateral force. Strain hardening will increase
the overstrength even further.
Moments induced in the cap-beam are also greatly influenced by the difference
between the moment pattems at first yield and full mechanism (solid and dashed lines in
Fig.3.25(d)), with the peak design moment increasing from 7.5 units to 16.5 units. Since
the cap-beam is designed to remain elastic, it is essential that this moment increase be
accounted for in design.
‘The second consideration relates to the column stiffness assumed for the gravity load
analysis, which as noted above will probably be the gross-section stiffness. During design-
level seismic response, the effective stiffness of the columns will reduce to about 30-30%
of the gross stiffness (see Section 4.4) when the column reaches first yield, and to perhaps
10% or less of the gross-section stiffness at maximum displacement response. As a
consequence of the reduced column stiffness, the gravity load moments in the columns
will almost entirely dissipate during the seismic response, as suggested by the dashed line
in Fig, 3.26(b). The question should be asked as to whether the values of gravity moments
are more relevant at the start of the seismic response than at the maximum response
displacement. Certainly the increase in positive gravity load bending moment in the cap
beam from 3 to 9, which can be expected to remain after seismic response must be
considered in design
‘The enhanced lateral strength of the structure indicates that the overall displacement
ductility demand will be proportionately less than intended. Although it may be argued
that this is necessary to avoid excessive curvature ductility demand, at the critical section
A.at the top of the column, it is in fact easy to show that this is a fallacy. As noted above,
we assume that the “equal displacement” approximation holds, and hence the expected
displacement ductility is 4 = R. The curvature ductility demand at A corresponding to
Hs is Hy The combination of moments and of curvatures is shown in Fig.3.26(e), which
relates to the moment-curvature response of section A, with a maximum design curvature
Of Omar = fg» The moment corresponding to the reduced seismic force is less than the
nominal moment capacity (60% of capacity, in the above example), and hence, assuming
a bilinear moment-curvature response, as shown ia Fig.3.26(¢), the seismic curvature is
less than the yield curvature by an equal amount, This is indicated in Fig.3.26(e) as
Peyeduced. The curvature demand will thus not be Qnas, since only the seismic componentChapter 3. Direct Dispiacement-Based Design: Fundamental Considerations 119
of the yield curvature will be increased by the expected curvature ductility factor to give
Ho Grretuced, 100 Hy.Gy, a8 shown in Fig.3.26(@). The curvature corresponding to the
gravity moment will be added to this, giving the final curvature as Pus
Fig.3.26(€). Thus the full expected design curvature at A has not been used,
Ic is best to illustrate this numerically. We assume that the design force reduction
factor and displacement ductility are R = fly = 5, and the corresponding design curvature
ductility demand is 44 = 10. Using the numerical values for gravity and seismic moment
in Fig. 3.26, the scismic curvature corresponding to the reduced seismic force iS Preduced
0.69. The plastic seismic curvature will be (4g - 1) Przreduced =90.6 9 =5.49,.
The total curvature will thus be ¢ + 5.4@, = 6.4@, and the curvature ductility demand,
at 6.4, is 64% of the design value. Thus even the most critical of the sections is subjected
10a much lower ductility demand than intended in che design.
On the basis of the above arguments, it is clear that direct addition of gravity and
seduced seismic moments is illogical, and unnecessarily conservative. A logical
improvement for force-based design would be to recognize that it is the sum of the
gzavity and unreduced seismic curvatures which should equate to the design curvature
mit, It follows from this thar the gravity and seismic moments should be combined
according to:
as shown in
(Mg + M,)/RS My (3.50)
where My is the nominal moment capacity of the section. The gravity moments should
be calculated using stiffness values that are compatible with those used for the seismic
analyses. This approach greatly reduces the influence of gravity moments in seismic
design, and in many cases gravity moments will become insignificant.
3.7.2 Combination of Gravity and Seismic Moments in Displacement-Based
Design
The arguments developed above become more critical when related to direct
displacement-based design. Because of the great differences in effective stiffness used to
determine gravity moments (gross section stiffness) and seismic moments (eracked-
section stiffness reduced by ductility factor — see Section 4.4) resulting moment
combinations would be meaningless. The appropriate combination must be consistent
with the design philosophy that we are concerned with conditions at maximum
displacement response, not in the clastic state. Therefore, the gravity moments should be
determined using the same effective stiffness as appropriate for the seismic design. In the
example above, this would mean using greatly reduced elastic stiffness for the columns,
swith the end results that gravity moments would become almost insignificant,
It should be noted, however, that the reason that the above approach is valid, is that
she gravity moments in the columns are based on compatibility, rather than equilibrium
requirements, Care is needed to ensure that structures, or parts of structures where
etavity moments are based on equilibrium considerations, do not have thet moments121
Chapter 3. Direct Displacement-Based Design: Fundamental Considera
ne = has Dk : re = Dkyz, hy (
) stiffness in the Z and X directions
where ky and ky are element (ie. walls or fram.
respectively, and x; and 4 are measured from the centre of ma
hee
Cu watt
Zz Wat
Was
(a) Structural Wall Building
(©) Frame with Eccentric Service Core
Fig.3.27 Examples of Structures Asymmetric in PlanChapter 3. Dicect Displacement-Based Design: Fundamental Considerations 123
Fig.3.28 Torsional Response of an Asymmetsic Wall Building (Plan View)
Note thar the strength eccentricity is used in Eq.(3.53). ‘The vwist angle @ to be
considered in Eq(3.53) is found from the total building strength in the direction
considered, Vgz, and the effective rotational stiffness Jeg as
ORV gz Cay IS, (3.55)
taf
where
Jag =F Mikal env) + Dhawles eee (3.56)
Note that the stiffhess eccentricity is used in Eq.(3.56), and the elastic stiffness of
clements responding in the Z direction is reduced by the design system ductility, ys
Since the transverse (K direction) elements are expected to remain clastic, or nearly
clastic, theie elastic stiffness is not reduced. Thus, with reference again to Fig.3.28 under
Z direction excitation:
Jroy = Ko s(O5Ly —eal) +h aa(0-5Ly +egl) Velen +2has(O5Lz¥ G57)
Note that the stiffness eccentricity egy in Fig.(3.55) can be based on ether elastic or
effective stiffness. Since the latter are each taken as ket/ flys, identical values are obtained.
In Eq.(3.57), itis assumed that the elastic stiffness of walls 3 and 4 is the same.
Since torsional eccentricity is most common in wall buildings the justification for the
above approach, and further discussion of torsional effects, is presented in Chapter 6,124 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
3.8.3 Design to Include Torsional Effects
(a) Design to Avoid Strength Eccentricity: The first objective in consideration of
torsion will be to investigate the possibility of eliminating the problem. With a building
whose plan layout is symmetrical in location and in dimensions of lateral force-resisting
elements this is of course no problem. Even when the plan layout of lateral force resisting
elements is unsymmetrical, it may be possible to eliminate strength eccentricity. With
respect to Fig.3.28, it may be possible to assign equal strengths to the two end walls (1
and 2) despite their different lengths by increasing the flexural reinforcement content of
wall 2 in comparison with that of wall 1. This would occur during allocation of the total
base shear between the different elements (Section 3.5.6). Note that the approach
outlined in the previous section indicates that there will stil be a torsional component of
response despite the zero strength eccentricity, which needs to be considered in the
design process outlined in (c) below. Torsional response of systems with stiffness
eccentricity, but no strength eccentricity is confirmed by inelastic time-history results
(b) Design to Minimize Strength Eccentricity: When the plan layout of lateral force-
resisting elements is such that strength eccentricity is unavoidable, the design objective
will generally be to minimize the strength eccentricity, so that the inelastic twist will be
minimized. This will be the case when the drift of the most flexible element governs
displacement-based design, but may not be the optimum solution when ductility capacity
of the stiffest element governs (see Section 6.4.5).
(©) Modification of Design Displacement to Account for Torsion: ‘the most
common design situation, particularly for frame buildings, bat also for many wall
buildings of more than four storeys, will be that design displacements are governed by
code drift limits. In these cases, the code drift will apply to the element with greatest
displacement, including torsional effects, meaning that the design displacement at the
building centre of mass, used in the SDOF design, will need to be reduced in proportion
to the torsional displacements. The design displacement for the centre of mass will thus
be found, reorganizing Eq.(3.53) to give:
Boy = Ai ~ 8, — ev) (3.58)
where Aj is the drift-controlled displacement of the critical element. With reference
again to Fig.3.28, and assuming that drift limits apply to wall 2, the design displacement
for the SDOF substitute structure will be
Key = Az — A(0-SLy + ey) (3.59)Chapter 3. Direct Displacement-Based Design: Fundamental Considerations 125
Icis also possible, particularly for low-rise wall buildings, or buildings containing walls
with low height/length aspect ratios, that the displacement capacity of the stiffest wall
corresponding to material strain limits may govern design. In this case the design
displacement at the centre of mass will be larger than the displacement of the critical
element, Equation (3.56) still applies, with due consideration of signs, and with reference
to Fig,3.28, the design displacement for the substitute structure will be
Bey =A, + (0.5L, —le,x|) (3.60)
In general it will be necessary to adopt an iterative approach to determine the design
displacement when torsional effects are significant, since @ depends on Je, eg and ev
which in turn depend on the relative strengths assigned to the lateral force-resisting
clements in both orthogonal directions, and the system ductility. However, adequate
simplifying assumptions can often be made to avoid the necessity for iteration. Since this,
is mainly relevant to the behaviour of wall structures, it is discussed in further detail in
Chapter 6.
3.9 CAPACITY DESIGN FOR DIRECT DISPLACEMENT-BASED DESIGN
Direct displacement-based design is a procedure for determining the required strength
of different structural systems to ensure that a given performance state, defined by
flexural strain or drift limits, is achieved under a specified level of seismic intensity, From
this design strength, the required moment capacity at intended locations of plastic hinges
or shear capacity of seismic isolation devices, with seismic isolated structures) can be
determined. As with force-based design, it is essential to ensure that inelastic action
occurs only in these intended locations, and only in the desired inelastic mode. For
example, a cantilever wall building will have intended plastic hinges at the bases of the
various walls, where inelastic action will be required to occur by inelastic flexural rotation.
Special measures are required to ensure that unintended plastic hinges do not occur at
other locations up che wall height, where adequate detailing for ductility has not been
provided, and to ensure that inelastic shear displacements, which are accompanied by
rapid strength degradation, do not occur.
Moments and shears throughout the structure resulting from the distribution of the
base shear in accordance with Sections 3.5.6 and 3.5.7 inchude only the effects of the first
inelastic mode of vibration. This is adequate for determining the required strength at
plastic hinge locations. However, actual response of the structure will include effects of
higher modes, These will not affect the moments at the plastic hinge locations, as these
are defined by, and limited to, the first inelastic mode values, but will influence moments
and shears at other locations.
A further factor to be considered is that conservative estimates of material strengths
will normally be adopted when determining the size of members, and (for reinforced
concrete design) the amount of reinforcing steel. If the material strengths exceed the
design values, as will normally be the case, then the moments developed at the plasticChapter 3. Direct Displacement-Based Design: Fundamental Considerations 127
higher mode, and may amplify the gravity moments considerably. A strict formulation of
capacity design would take this into account.
Column end moments and shear forces are amplified for both beam plastic hinge
overstrength and dynamic amplification. For one-way frames, upper limits for dynamic
amplification of column moments of 1.80 have been recommended, with 1.3 for column
shear forces. Further amplification for beam flexural overstrength is required",
In this section we have discussed conventional capacity design, as currently applied to
force-based design of structures. We show in Section 4,5 and the design chapters related
10 specific structural types that modifications to the capacity factors for both flexural
overstrength, and dynamic amplification are appropriate for direct displacement-based
design.
1
_ 08
& &
= o6 5 “S._+ Dynamic
4 3 *Ampifiation
5 g
2 04 z
z z
a a
02
0 i
oO 04 08 1.2 0 0.5 1 15 2 25
Dimensionless Moment Dimensionless Shear Force
(a) Moment Profiles (b) Shear Force Profiles
Fig. 3.29 Recent Recommendations for Dynamic Amplification of Design
Forces for Equivalent Lateral Force Design of Cantilever Walls!”!1
3.10 SOME IMPLICATIONS OF DDBD.
3.10.1 Influence of Seismic Intensity on Design Base Shear Strength
Direct displacement-based seismic design implies significantly different structural
sensitivity to seismic intensity than found from current codified force-based design
procedures. This can be illustrated with reference to Fig.3.30, where acceleration spectra
(Fi.3.30(@)), and displacement spectra (Fig.3.30(b)) are shown for two seismic zones. It128 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
is assumed that the spectral shapes for the two zones are identical, and each are found by
multiplying a base-leve! spectrum by the zone factors Z1 or Z2.
Consider the design of two reinfoxced concrete buildings, one designed for each of the
seismic zones, where the structural geometry, including member sizes (but not
reinforcement contents) are identical for the two buildings, We assume that structures are
designed to exactly satisfy the strength requirements for the two zones. If the buildings
are designed by conventional force-based procedures, the fundamental periods of the two
buildings will be assumed to be the same, since the same allowance for reduction of
gtoss-section stiffness will be made. Assuming the same force-reduction factor is used for
each design it is thus clear (see Fig.3.30(a)) that the required base-shear design forces Vs
and V2 for the two buildings are related by:
Z,
(3.63)
"Z,
(a) Acceleration Spectra (b) Displacement Spectra
Fig.3.30 Influence of Seismic Intensity on Design Base Shear Force
Under direct displacement-based design, the assumption of equal geometry ensures
that the yield displacements, and the limit-state design displacements, based on drift
limits, for the two buildings are the same. Hence the ductility, and also the effective
damping will also be the same for the two buildings, As may be seen from Fig.3.30(b),
with equal design displacements and damping, the effective periods at design
displacement response will be related to the zone intensity byChapter 3. Direct Displacement-Based Design: Fundamental Considerations 129
From Eq,G.1) the required effective stiffness is inversely proportional to the period
squared, hence:
(3.65)
Further, since the design displacements are equal, F.q.(3.2) yields the ratio of base shear
forces as:
Vioses = 2 (3.66)
1)
‘Thus the required base shear strength is proportional to the square of the seismic
intensity. This is a fundamentally important difference between the two approaches,
particularly for regions of low (or very high) seismicity. It should be noted however, that
the difference between the conclusions resulting from force-based or displacement-based
considerations is largely a consequence of the assumption in the force-based approach
that stiffness may be assumed to be equal to a constant fraction of the gross-section
stiffness. As has been pointed out in relation to Fig.1.4, and is discussed in some detail in
Section 4.4, stiffness of a member is directly proportional to strength. Thus the elastic
stiffness of the members of the structure in the lower seismic zone will be less than for
the structure in the higher seismic zone, An iterative force-based solution, correcting the
stiffness of the structural members based on the strength found from the previcus
iteration would eventually come to a similat conclusion to that resulting from DDBD.
3.10.2. Influence of Building Height on Required Frame Base Shear Strength
A further finding of some interest can be obtained by examining the sensitivity of
required base shear strength of buildings with identical plan geometry and storey mass,
but with different numbers of storeys. We assume for simplicity that the section
dimensions of structural members are not affected by building height, and that the design
deflected shape is also independent of building height. Clearly this latter assumption will
become increasingly crude when large variations in building height are considered, but is
reasonable for frame buildings up to about 10 storeys, and for dual wall/frame buildings
up to 20 storeys. Figure 3.31 compares two frames of different heights. Let 2 = number
reys, with constant mass m per storey. In the following, C; to Cs are constants.
Effective mass: (3.672)
Design Displacement: (3.676)130 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Tw _————
(a) Building 1 (b) Building 2 (c) Design Displacements
Fig.3.31 Design Displacement Profiles of Buildings Differing only in Height
With the above assumptions, both the yield displacement, and the design displacement
will be proportional to height. Hence the design displacement ductility and thus the
effective damping will be independent of height, and the design displacement response of
the buildings will lie on the same damping curve (see Fig,3.1(d)). Provided that the design
displacement given by Eq,(3.67b) is less than the displacement at peak period (e.g. Te =4
sec, in Fig.3.1(d)), and that the design displacement spectrum is linear with period, the
effective period can thus be expressed as:
T= OAg=O,C,n G.67¢)
From Eq.(3.1), the effective stiffness will be:
om 6.67)
(CC)?
From Eqs.(3.2), (3.67b) and (3.674) the design base shear will be
Vie =A, =C.™ Cyn = C, 7
Voae = kg = Cs Cn = Cam (3.676)
n
Recalling that mis the mass of one storey, it is seen that the design base shear strength
is independeat of the number of storeys. This might seem to point the way towards
further possibie design simplifications
3.10.3 Bridge with Piers of Different Height
‘We return to the example of a bridge crossing a valley, with piers of different heights,
first discussed in Section 1.3.4, and subjected to longitudinal seismic excitation. Figure
1.11 from Section 1.3.4 is reproduced below as Fig,3,32(a).Chapter 3. Direct Displacement-Based Design: Fundamental Considerations 131
Displacement
(@) Structure (b) Force-Displacement Response
Fig.3.32 Bridge with Unequal Pier Heights under Longitudinal Excitation
As was discussed in Section 1.3.4, designing in strict accordance with force-based
design would require allocating sheat between the piers in proportion to the elastic
stiffnesses of the piers. This is based on the assumption that yield displacements, and
ductility demands for the piers can be equalized by distributing strength in proportion to
stiffness, We have showa this to be invalid, since the yield curvature, and hence the yield
displacement is essentially independent of strength. The consequences, as noted in Table
3.6 are that flexural reinforcement ratios for the piers should be approximately in
proportion to the inverse of pier height squared. ‘The sequence of design operations
follows the arrow in the second column of Table 3.6
With DDBD, the initial stiffness is largely irrelevant, and the relationship berween
flexural reinforcement ratios ia the piers is the designer’s choice. Normally equal
reinforcement ratios will be chosen, and the moment capacities of the flexural plastic
hinges will be essentially equal. This implies distributing the total seismic force benween
the piers in inverse proportion to height, as indicated in Table 3.6, Column 3, which also
shows that the sequence of design operations, indicated by the arrow, is the exact reverse
of that for force-based design. Of course it would be possible to use a rational
reinforcement distribution for force-based design, but only if the concept of the
importance of initial stiffness is abandoned.
‘Table 3.6 Difference between Force-based and Displacement-Based Design
Parameters for Bridge under Longitudinal Excitation
fr 1TEM Force-Based DDBD
Design
Yield Displacement Equal a
Ductility Demand Equal o.1/FP
Stiffness aif | at/H
Design Shear Force 1/HP aij
[Design Moment oe [Equal
Reinforcement Ratio Equal132 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
3.10.4 Building with Unequal Wall Lengths
The structural wall building shown in Fig.3.33, also considered in Chapter 1, and in
Design Examples 3.5 to 3,7 is now considered. Force-based design, relying on initial
stiffness distributes the design base shear force berween the different length walls in
proportion to wall length cubed on the invalid assumption that the walls can be made to
yield at the same displacement. Since the yield curvature is inversely proportional to wall
length, so will the yield displacements be. ‘Table 3.7 compares the proportions of the
design parameters based on initial-stiffness force-based design, and direct displacement-
based design.
A B c
Force
—
>
>
>
Co
1 Anos Ay Aaa
lek be heh lc Aye” Displacement
Fig.3.33 Building with Unequal Length Cantilever Walls
Table 3.7 Comparison of Design Parameters for Force-Based and DDBD Designs
of a Cantilever Wall Building
ITEM Force-Based | DDBD
Design
Yield Displacement Equal
Duetility Demand Equal
Stiffness ody”
Design Shear Force ole
Design Moment ale
‘Reinforcement Ratio Oh4
ANALYSIS TOOLS FOR DIRECT DISPLACEMENT-BASED
DESIGN
41 INTRODUCTION
It is beyond the scope of this book to include a detailed treatment of material
properties, methods of analysis for determining section strengths, and general global
structural analysis methods, There are, however, a number of aspects of these topics
which are of special relevance for direct displacement based design (DDBD), particularly
related to the strength and deformation characteristics of reinforced concrete and
masonry sections. Hence this chapter provides a detailed examination of selected topics.
More general treatments are available in other sources (?1."4 to which the interested reader
1s referred.
4.2. FORCE-DISPLACEMENT RESPONSE OF REINFORCED CONCRETE
MEMBERS
(a) Structure (b) Deflection (ce) Moment-Curvature (A) Force-Displacement
Fig.4.1 Lateral Deformation of a Bridge Column
One of the most basic tools for DDBD is she moment-curvature analysis of reinforced
concrete and masonry sections. This is used to define section strengths, limir state
curvatures, and also the elastic stiffness. From these data, member and structure force-
133134 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
displacement characteristics can be directly generated, as discussed in Section 4.2.8, and
suggested in Fig.4.1 for a simple bridge pier. The treatment in the following relates
specifically to reinforced concrete sections, though the principles are identical for
reinforced masonry. In general, moment-curvature analysis of steel sections is un-
necessary, since the elastic stiffness is directly known. Generation of force-displacement
response of steel sections is briefly covered in Section 4.3
4.2.1. Moment-Curvature Analysis
centroidal
(a) Symmetrical section _(b) Strains (©) Stresses
n layers of rebar
Fig.4.2. Strains and Stresses in an Arbitrary Symmetrical R.C. Section
Figure 4.2 shows ao arbitrary symmetrical reinforced concrete section subjected to
bending, with the top of the section in compression. The normal assumptions for flexural
analysis are made:
© The strain profile is linear at all stages of loading up to ultimate (ie. the Navier-
Bernoulli “plane-sections” hypothesis holds).
© Steel strain and concrete strain at a given distance from the neutral axis (see Fig.
4.2) are identical (ie. perfect bond between steel and concrete exists)
© Concrete and reinforcement non-linear stress-strain relationships are known:
Sees = P (Egy) = PE ny) (4.1@)
Savy =P Egy) = PE) (410)
Concrete tension strength is ignored in the analysis.
© Axial force (if any) is applied at the section centroid.Chapter 4. Analysis Tools for Direct Displacement-Based Design 135
As discussed subsequently, different concrete stress-strain relationships will apply to
core and cover concrete. Note that the reason that concrete tension strength is ignored in
the analysis is that the section will be subjected to reversed loading under seismic attack.
Since the neutral axis will generally be on the compression side of the section centroid,
cracks under reversed loading will extend through the entire section. ‘That is, the
compression zone will occur in a location previously cracked under moment of the
opposite sign, and no tension capacity will exist after application of service loads, or
preliminary low-level seismic response. Further, the contribution of concrete tension
strength to flexural strength is normally negligible.
Let the axial force on the section be denoted by N. Then for axial force equilibrium at
any level of response, using the nomenclature of Fig 4.2:
N= [fo nd tS fod, = JO. (6) +L, (€y,)4, 42)
where Ay is the total area of all reinforcing bars at layer J, distance y; from the centroidal
tically, the position of the neutral axis is adjusted by trial and error until
Eq.(4.2) is satisfied
‘Taking moments of the stress resultants about the section centroid:
M= JOE vd LOE IVA, (43)
7
“The corresponding curvature is given by
=fe5_fu 4.4)
oe G-o a
where & and & respectively are the extreme-fibre compression strain, and strain at the
level of the reinforcing bars at maximum distance from the neutral axis (see Fig.4.2)..
Moment-curvature analysis is normally organized in accordance with the following
steps:
1. Divide the section into a number of stices perpendicular to the loading axis,
Determine the area of unconfined cover concrete, confined core concrete, and
reinforcing steel in each layer.
2. Select an extreme fibre compression strain, starting with the lowest value.
3. Assume a neutral axis location.
4. Calculate concrete and steel stresses at the centre of each layer, and hence the
concrete and steel forces in each layer
5. Check axial force equilibrium in accordance with Eq.(4.2).
6. Modify the neutral axis position to improve agreement in Eq.(4.2).136 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
7. Cycle steps 3 to 6 until satisfactory agreement is obtained between external and
internal forces.
Iculate the moment (Fg/4.3)) and curvature (Eq.(4.4)).
9. Increment the extreme fibre compression strain, and repeat steps 3 to 8.
10. Continue incrementing the extreme fibre compression strain until the ultimate
compression strain (discussed subsequently) is reached.
‘ 4.2.2. Concrete Properties for Moment-Curvature Analysis
: confined coneréte
hoop fracture”
Compressive Stress f.
>
ungonfined concrete
rs
Exp Exe
Compressive Strain, &
Fig.4.3 Stress-Strain Model for Concrete in Compression M41
As noted above, separate stress-strain relationships should be used for the
compression response of unconfined cover concrete and confined core concrete. As
shown in Fig. 4.3, confined concrete has increased compression strength, and more
importantly, increased compression strain capacity. The enhancement of compression
stress- strain characteristics of the core concrete is a result of the action of well-detailed
transverse reinforcement in the form of hoops or spirals. In conjunction with longitudinal
reinforcement, close-spaced transverse reinforcement acts to restrain the lateral
expansion of the concrete that accompanies the onset of crushing, maintaining the
integrity of the core concrete.
Figure 4.4 shows four column sections confined by different configurations of
transverse reinforcement. In Fig.4.4(a), the confinement is provided by circular spirals ot
hoops. As the concrete attempts to expand, it bears uniformly against the hoop or spiral,
placing it in tension, resulting in a uniform radial compression on the concrete.Chapter 4, Analysis Tools for Direct Displacement-Based Design 137
(a) Circular hoops —_(b) Rectangular hoops __(c) Rectangular octagonal
or spirals with cross-ties hoops
unconfined
concrete
Wl
() Overlapping rectangular (e) Confinement by _(f) Confinement by
hoops transverse rebar longitudinal rebar
Fig.4.4 Confinement of Column Sections by Transverse Hoops and Spirals!
In Fig4.4 the unconfined cover concrete is shown shaded. It is normally assumed
that the maximum effective radial confining pressure, fj that can be exerted on the core
concrete occurs when the spirals or hoops are stressed to their yield stress fry. ‘Taking a
free body across the column diameter, and noting that this exposes two bar forces fry
where Ap is the spiral or hoop cross-sectional area, the confining stress in the core is:
2huds
D's
=05—Son (45)
where sis the spacing of the hoop or spiral along the column longitudinal axis, D’is the
diameter of the confined core, measured to the centreline of the hoop or spiral, and p,
4A,/D"s is the volumetric ratio of transverse hoops or spirals.
Note that the longitudinal spacing s of the hoops or spiral also affects confinement
efficiency, as shown in Fig, 4.4(¢), since the confining effect is concentrated as a line load
at the level of the spiral. However, this load is partially distributed to the longitudinal
reinforcement which tends to make the confining effect more uniform (see Fig.4.4(6),
illustrating the importance of the longitudinal reinforcement to confinement of concrete.138 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
The hoops or spirals also act to restrain the longitudinal reinforcement from buckling
when in compression. Buckling may occur after longitudinal reinforcement is. first
subjected to inelastic tension strain under one direction of seismic loading. When the
loading direction is reversed, the bars initially transfer all the compression force on the
section, and must yield in compression before previously formed cracks close. It is during
this stage of response that the bars are susceptible to buckling. Once the cracks close, the
compression stiffness of the concrete can be expected to restrain the tendency for bar
buckling, It will thus be seen that bar buckling is more dependent on the inelastic tensile
strain developed in a previous yield excursion, than on pure compression characteristics
To ensure inelastic buckling does not occur, the maximum spacing of transverse hoops or
spirals must be related to the bar diameter, as discussed further in Section 4.2.5(c)
With rectangular sections, it is possible to develop different levels of lateral confining
stress in orthogonal directions. If Ayu is the total amount of lateral reinforcement in a
hoop layer crossing a section perpendicular t0 Ay4x, then the maximum confining stress
that can be developed in that direction is
CAatcbri
hys
ty
Sic = = OP aS, (4.6)
where Pax is the area ratio of transverse reinforcement in the x direction, is again the
spacing of hoop sets along the member axis, and €, is a confinement effectiveness
coefficient, relating the minimum area of the effectively confined core (see Fig.4.4) to the
nominal core area bounded by the centreline of the peripheral hoop, and fy is the core
width perpendicular to the direction considered, measured to the centreline of the
petipheral hoop.
Clearly a similar expression can be developed in the orthogonal direction. Ideally,
rectangular sections should be designed to have equal area ratios of confining
reinforcement in the orthogonal directions, in which case the volumetric confinement
ratio is:
P.
Px + Pay =2Po (4.7)
and the lateral confining stress is given by
f= 95CP.fy 48)
The similarity to Eq is obvious. Note that a confining effectiveness coefficient C. is
sometimes also assigned to circular sections, but since it will normally be higher than 0.95
for typical designs, itis generally taken as 1.0. For rectangular sections, values of between
0.75 and 0.85 are appropriate, depending on the ratio s/fe, and the number of
longitudinal bars in the section. For walls, C, = 0.5 is generally appropriate.140 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
4.2.4 Reinforcing Steel Properties for Moment-Curvature Analyses
compression strain tension strain
Fig. 4.5 Reinforcing Steel Stress-Strain Characteristics
In conventional seismic design the rcinforcing steel design strength is often taken as
the yield strength. That is, the increase in stress due to strain-hardening is generally
ignored. When carrying out moment-curvature analyses however, it is important that the
most realistic representation of the full stress-strain characteristic be utilized, including
the strain-hardening portion. ‘This is particularly important for DDBD, because the
design attempts to match the strength at the expected design displacement to the required
strength, rather than to use an artificially low nominal “yield” strength.
Figure 4.3 illustrates che characteristics of typical monotonic and cyclic stress-strain
response of reinforcing steel, The monotonic response can be represented by three
phases:
Elastic: 056, $6: S.=E£,81f, (4.16)
Yield plateau: 65656, f, = f, (4.17)
Strain-hardening: &45656.: f,~ f,)1-(f,-f, fé é ] (4.18)
\ Em ~ Eh
In Fg(4.16) to (4.18), & and f, and the reinforcing steel strain and stress, Ey is the
clastic modulus, and the other symbols are defined in Fig.4.4. Note that the curve only
continues up to Ew, the strain at ultimare stress, as behaviour past this point isChapter 4, Analysis Tools for Direct Displacement-Based Design 141
characterized by local necking of the reinforcement, with reduction of stress elsewhere,
and cannot be relied on in design,
Typical values for the characteristic stresses and strains depend on the type, and grade
of reinforcement used. In seismic applications in the American continent, ASTM 7068
grade 60 reinforcement (f, minimum = 414MPa) is normally used. Strain-hardening
usually starts at about &» = 0.008, the ultimate strain is about &y = 0.10 to 0.12, and the
ratio of ultimate to yield stress is typically f/f, = 1.35 to 1.50. In Europe tempcore
reinforcement is almost always used, with a higher yield stress (typically f, = 500MPa
(72.5ksi)), an ulkimate strain of about 0.09, a ratio of ultimate to yield stress of about fulfy
= 1.2, and essentially no pronounced yield plateau.
Asis illustrated in Fig. 4.5, the cyclic characteristics of the reinforcing steel differ from
the monotonic curve, On unloading and reloading, the stress-strain curve softens early,
due 0 the Bauschinger effect, and no pronounced yield plateau is apparent. However, it
is found that force-displacement response predicted from moment-curvature
characteristic based on the monotonic stress-strain curves for both concrete and
reinforcement provide a good envelope to measured cyclic response, Where full cyclic
moment-curvature response is required (this is more normally the case for research
applications rather than for design), more complete equations are appropriate, and the
reader is referred elsewhere!'I,
4.2.5 Strain Limits for Moment-Curvature Analysis
(2) Damage-Control Compression Strain: The useful limit to confined concrete
compression strain is usually taken to occur when fracture of the transverse
zeinforcement confining the core occurs. This may be estimated by equating the increase
in strain-energy absorbed by the concrete, above the value appropriate for unconfined
concrete to the strain-energy capacity of the confining steel. Considering a unit volume of
core concrete under uniform axial compression, the increase in strain-energy absorbed
can be expressed as:
SEcone = OS ce Een (4.19)
where Cy is a coefficient dependent on the shape of the unconfined and confined stress
strain curves. Similarly, the strain-energy capacity of the transverse reinforcement, related
£0 the same unit volume of core concrete can be expressed as
SE tar = CoP SE wn (4.20)
where Cz depends on the shape of the reinforcing steel stress-strain cufve, ‘Thus equating
4.19) and (4,20), and assuming that the unconfined ultimate strain of the concrete is
0,004, the following expression for the ultimate compression strain for confined concrete
can be derived:142 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
ahh
fre
= 0.004 +1 (4.21)
In Eq.(4.21) an average value of G/G=1.4 is adopted, based on typical steel and
concrete stress-strain shapes. Equation (4.21) is approximate for a number of reasons:
‘Iris based on pure axial compression of the core concrete. Under combined axial
compression and flexure, the equation will predict conservatively low estimates
for Eu
© Ir does not take into account the additional confinement of the critical section by
the adjacent member. If the member is a column supported on a foundation,
then the foundation will act to restrain lateral expansion of the critical lower part
of the column plastic hinge, thus adding to the confining effect. Similarly, the
end portion of the plastic hinge of a beam framing into a column will be
confined to some extent by the column, provided the column remains elastic.
Although experimentsil have shown the equation to provide good estimates of
ultimate compression strain, the logic (energy balance) does not appear « apply
at reduced levels of extreme fibre compression strain. At these levels, inverting
Eq.(4.21) to provide estimates of transverse reinforcement strain (6, rather than
&,) over-predicts the measured strains
The combined effects of these approximations results in effective ultimate
compression strains under combined axial force and flexure that exceed the predicted
values by a factor of about 1.3 10 1.6. The influence on ultimate displacement is similar.
Our view is that this degree of conservatism is appropriate for structures designed t0 a
“damage control” limit state (see Section 3.3.2).
(®) Serviceability Limit Compression Strain: The limit compression strain for
concrete corresponding to the serviceability limic state should he a conservative estimate
of the strain at which spalling initiates, Below this strain-limit repair should not be needed
which is compatible with performance criteria for the serviceability limit state. In seismic
response, maximum compression strains almost always occur adjacent to a supporting,
member (c.g, a foundation beam, for a concrete column or wall), which provide an
additional restraint against initiation of spalling. Experiments!” have indicated that a
compression strain of Es = 0,004 is a conservative lower limit to initiation of spalling,
and this value will be used for concrete structures in this text. For masonry structures a
somewhat more conservative estimate is appropriate, and we recommend &. = 0.003.
(©) Damage-Control Tension Strain Limit: \t is inappropriate to use Em , the strain
at _maximum stress of the reinforcing steel found from monotonic testing, as the
maximum permissible tension strain for moment-curvature analysis for several reasons.
First, under cyclic loading, the effective ultimate tensile strain is reduced by the peak
compressive strain, &, achieved under a previous reversal of loading direction, asChapter 4. Analysis Tools for Direct Displacement-Based Design 143
suggested in Fig. 4.419, Second, once high tensile strains have been developed, the
longitudinal reinforcement becomes susceptible to buckling when subjected to reversed
ioading that puts the reinforcement ir. compression. This buckling typically occurs before
the previously developed flexural cracks are closed, and while the bars are still subject t©
rensile strain (but compressive stress). This engenders low-cycle fatigue of the reinforcing
bar at levels of tensile strain significantly below &,. The level of strain will depend on the
volumetric ratio and longitudinal spacing of the transverse reinforcement. Finally, slip
between reinforcing steel and concrete at the critical section, and tension-shift effects
result in reinforcement strain levels being lower than predicted by a “planc-sections”
hypothesis. Based oo these considerations, the ultimate curvature of the section analysed
should be based on a steel tension strain limit of & = 0.6€s», if this occurs before the
concrete ultimate compression strain £4, is developed. To ensure this level of strain is
attainable without the reinforcement buckling, the spacing of transverse reinforcement
hoops or layers should not exceed s = (3 + 6(f/fy-1)) dyy (dy = longitudinal bar diameter)
(d) Serviceability Limit Tension Strain: \t has bees common in the past to require
“elastic” or “near-elastic” response at the serviceability limit state, This is generally taken
to mean that the displacement ductility demand should be /£ © 1, implying reinforcement
tension strains that are at, or only slightly above yield strain. ‘This ‘s, in our view
excessively conservative, as strains of several times the yield strain can be sustained
without creating damage requiring repair. The critical aspect is likely t0 be the crack
widths developed by the seismic response, Moreover, it is not the instantaneous
maximum value occurring during seismic response, but the residual crack width that will
be of concern, since potential corrosion is the issue here. Analyses ceported elsewhere!
indicate that for structural elements with compression gravity loading (walls, columns), a
maximum tension strain of 0.015 during seismic response will correspond to residual
crack widths of about 1.0 mm (0.04in), and for members without axial compression (e.g,
beams) a peak strain of about 0,010 is appropriate. A residual ctack width of 1.0mm
should not need remedial action in normal environments. In aggressively corrosive
environments, lower residual crack widths may be appropriate, with correspondingly
reduced serviceability tensile steain limits.
4.2.6 Material Design Strengths for Direct Displacement-Based Design
In gravity-load design of structures, it is common practice t0 assume minimum, or
characteristic lower bound values for material strengths when determining the nominal
strength of sections. This is typically combined with a strength reduction factor (or with
partial material factors) to ensure a conservative estimate of the section strength. Thus,
the design flexural strength requirement may be expressed as:
o,My>M, (4.22)144 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Susetures
where @ is the flexural strength reduction factor (typically gy $0.9), and My and Mp are
the nominal (designed) and required moment capacities.
‘This conservative approach has obvious relevance for gravity-load design, where the
consequences of the dependable strength (gy My) being less than the moment demand
could be catastrophic failure. However, in seismic design, it is expected that under design
seismic attack, the moment capacity will be considerably less than the demand resulting
from fully clastic response. Consequently, incorporation of extreme conservatism in
estimates of matetial strengths, and use of strength reduction factors will not result in
“protecting” the section against inelastic action. All that will happen is that the section
strength will be higher than needed, but will still be developed, with considerable ductility
demand, under design seismic attack. A small reduction in displacement demand might
result, buc this is best directly incorporated into the specified design displacement.
Consequently it is recommended that flexural strength reduction factors not be used
when designing locations of intended plastic hinging. Based on recommendations for
seismic design of bridges(4l ir is also recommended that the following design material
strengths be adopted:
Concrete: fee = N.3f. (4.232)
Steel fe = Af, (4.23b)
In Fq.(4.23) free and fre are low estimates of expected strength. The value for fre is felt
to be appropriate for both zeinforcing and structural steel. The concrete strength
acknowledges the influence of conservative batching practice (average 28-day
compression strength is typically 20% above the specified value), and the increase in
strength after 28 days before the structure is subjected to the design loading, which will
certainly occur later than 28 days.
Asis discussed in Section 4.5, it will also be necessary to obtain estimates of maximum
feasible strength in plastic hinges, for capacity design calculations, using upper-bound
material strengths. The following values are recommended:
Concrete: freo= \.1fc (4.24a)
Steel: fy =13fy (4.24b)
4.2.7. Bilinear Idealization of Concrete Moment-Curvature Curves
For design purposes, it is generally of sufficient accuracy to use a bilinear
approximation to che moment-curvature response, consisting of an initial “elastic”
branch, and a post-yield “plastic” branch. For reinforced concrete and masonry sections,
it is important that the elastic branch not be based on the initial uncracked section
stiffness, as this value is only appropriate for very low levels of seismic response. The
normal procedure is to use the secant stiffness from the origin through first yield as the
effective elastic stiffness. First yield is defined as the point on the moment-curvatureChapter 4. Analysis Tools for Dicect Displacement-Based Design. 145
response when the extreme tension reinforcement (ic. the rebar furthest from the neutral
axis) first attains yield strain, or when the extreme concrete compression fibre (again, at
‘maximum distance from the neutral axis) attains a strain of 0.002, which ever occurs first.
The moment and curvature at first yield are denoted M, and 9, respectively. This line
defining the elastic stiffness is extrapolated up to the nominal moment capacity, which is
defined by an extreme fibre compression strain of 0.004 or an extreme tensiva
reinforcing bar strain of 0.015, which ever occurs first. The corresponding curvature is
termed the nominal yield curvature 4,
The plastic branch is defined by connecting the nominal yield point (My, @) to the
ultimate condition: Mu, Gy.
This procedure is illustrated for a rectangular column section in Fig. 4.6, where the full
moment-curvature relationship is shown in Fig-4.6(a), and the initial portion is shown in
Fig.4.6(b) to expand the elastic branch, From Fig. 4.6(b) it is seen that the nominal yield
curvature is defined as:
(4.25)
The elastic stiffness is the slope of the initial branch. That is:
M, M.
EL, == (4.26)
go,
and the stiffness of the plastic branch is given by:
El, = aM (427
9-9,
The column represented in Fig, 4.6 had a square section, 800x800mm (31.5x31.5 in),
had a longitudinal reinforcement ratio of 1.88%, and a transverse reinforcement area ratio
of 0.436% (equivalent to a volumetric ratio of 0.872%). Material properties were f'e =
30MPa (4350 psi), f= fon = 425 MPa (61,600 psi, and axial load = 2MN (450 kips)..
In Fig. 4.6(b), the cracking moment Mg, and curvature $,, have also been identified.
For reasons elaborated above in section 4.2.1 these have been based on the assumption
of zeto tension strength for the concrete, Note that the elastic stiffness defined by
Eg.(4.26) is only about 40% of the initial stiffness of the un-cracked section.146 Priestley, Calvi and Kowalsky, Displacement-Based Seismic Design of Structures
3000 5
MyF>
2000 |
&
&
z : '
z i i
1000 + El '
} ' i
tore
0 0.02 0.04 0.06 0.08 “oa
Curvature (en)
(2) Full Moment-curvature response
2500 4
My
1500 4
Moment (kN)
g
L
0.01 0.015 0.02 0.025
Curvature (mi)
(6) Initial section Of Moment-Curvature Response
Fig.4.6 Example Moment-Curvature Curve for an 800x800 mm (31.5%31.5 in)
Column SectionChapter 4. Analysis Tools for Dircet Displacement-Based Design 147
4.2.8 Force-Displacement Response from Moment-Curvature
A
Fr, wee
a
H
May An)
1 _ -
a erertnaaccelan
(a) Structure (b) Moments (c) Curvature (4) Displacement
Fig.4.7. Obtaining Displacements from Curvature Distributions
Figure 4.7 illustrates what might appear to be the obvious method for obtaining the
placement at the top of a simple bridge pier subjected to a specified lateral force F,
‘om the moment-curvature relationship, and to hence build the force-displacement
ponse. The procedure, though related to a bridge pier for simplicity, is general for
slumns and walls, and can be easily adapted to develop force-rotation response, which
ay be more general for (say) beam members, Using the nomenclature of Fig4.7, and
measuring the distance ft down from the line of application of the inertia force, the
cnoment at h and at the base ( =H) will be given by:
My, =F-h; and M,=F-H (4.28)
The curvatures at all heights # could then be read from the moment-curvature
-elationship to produce the curvature distribution gj) shown in Fig. 4.7(¢), which could
be integrated to provide the top displacement, A, as:
H,
A= [Ohh (4.29)
a
Repeating this process for values of 0S FE M,/H would then be expected to provide
he full force-displacement response.
Unfortunately this process does not produce force-displacement predictions that agree
well with experimental results. These are a number of reasons for this:148
Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Tension shift is ignored. As shown in Fig-4,7(a), the influence of shear force is
likely to incline the flexural cracks from the horizontal orientation appropriate
for pure flexure. The result is that the “plane-sections” hypothesis is incorrect,
and reinforcement tension strains at a given level will be higher than predicted
from the moment at that sectionl"'l, Thus a simple integration of the theoretical
curvature distribution is an approximation
The approach ignores shear deformation, though this can be separately added to
the displacement response (see Section 4.8, e.g.)
Anchorage deformation (strain-penetration) is ignored. Equation (4.29) implies
that the curvature drops to zero immediately below the column base (or, for a
beam, at the column face). In fact, strains of the tension reinforcement will only
drop to zero at a depth equal to the true development length of the
reinforcement. This implies a partial pullout of the bar at the column base
section, which can be estimated by integrating the reinforcement strain profile
below the base. On the other side of the column, the concrete compression
strains will also not immediately drop to zero at the base, but will gradually
dissipate with depth. The reinforcement tension strain effect is more important,
and it is useful to define a “strain penetration” length Lsp, over which the
curvature may be considered constant and equal to the column base curvature,
In some cases the moment-curvature response will exhibit negative stiffness; that
is, the moment will decrease as the curvature increases. This will generally be the
result of reduced effective section size caused by spalling of cover concrete.
Attempting to integrate the curvature distribution in accordance with Eq(4.29)
will result in only one section (the base) of infinitesimal length, being subjected
to increased curvature as the moment drops. All other sections will not have
reached the peak moment, and hence will presumably have a reduction of
curvature. Strict application of Eq,(4.29) would then result in the ultimate
displacement being obtained as soon as the critical section reached the peak
moment. This does not accord with observations thar che displacement
continues to increase as the moment decreases
The solution to these problems is to use a simplified approach based on the concept
of a “plastic hinge”, of length Lp, over which strain and curvature are considered to be
equal to the maximum value at the column base. The plastic hinge length incorporates the
strain penetration length Lsp as shown in Fig.4.8. Further, the curvature distribution
higher up the column is assumed to be linear, in accordance with the bilinear
approximation to the moment-curvatute response. This tends to compensate for the
increase in displacement resulting from tension shift, and, at least partially, for shear
deformation
‘The strain penetration length, Lsp may be taken as:
Lp = 0.022 f,.d; (fre in MPa);
15 feds) Gye in ksi) (4.30)Chapter 4. Analysis Tools for Direct Displacement-Based Design 149
linear to
yield
I
Atego
Fig.4.8 Idealizarion of Cusvature Distribution
where fre and dy are the yield strength and diameter of the longitudinal reinforcement,
nd the plastic hinge length, Lp, for beams and columns, is given by:
Ly = ke + Lop 2 2Lgp (4.312)
where
ff ] £0.08 (4.31b)
Uy
and where Lc is the length from. the critical section to the point of contraflexure in the
member. Equation (4.31b) emphasises the importance of the ratio of ultimate tensile
strength to yield strength of the flexural reinforcement. If this value is high, plastic
Geformations spread away from the critical section as the feinforcement at the critical
section strain-hardens, increasing the plastic hinge length. If the reinforcing steel has a
‘ow ratio of ultimate to yield strength, plasticity concentrates close to the critical section,
zesulting in a short plastic hinge length. A modification of Eq(4.31) for walls is suggested
‘a Section 6.2.1(b)..
In Figs. 4.7 and 4.8, Le = H. The lower limit of Lp = 2Esp implies strain penetration
soth down into the foundation, and also up into the column, and applies when Le is
short. The force displacement response, for the cantilever column of Fig.4.7 can then be
mbled from the moment-curvatute response using the following equations:150 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
F=M/IH (4.32)
A, =9,(H +L) /3 (4.33)
A, =A, +A, =A, +0,LpH =A, +(9,-$, Lol 434)
Equation (4.34) implies that the centre of plastic rotation occurs at the member end.
This will be exact when Lp =2Lsp, and is an acceptable approximation in all cases.
However, when 0.08L¢ > Lsp, an improved estimate of the plastic displacement can be
obtained by replacing Hin Eq.(4.34) by the distance from the centre of the plastic hinge
to the point of contraflexure for members in single bending, and by the centre to centre
distance of plastic hinges at member ends for members in double bending.
bilinear
Displacement Ae
Fig.4.9 Force-Displacement Response
The resulting bilinear force-displacement response is shown as the dash-dot line in
Fig.4.9. This is normally adequate for design purposes. However, a more accurate
representation is possible, as shown by the solid line in Fig4.9. In this “refined”
approach, suitable for prediction of experimental response, the clastic portion is
represented by a bilinear characteristic, with the cracking force and displacement at the
corner, joined to the force and displacement at first yield, Above first yield, the
displacement is based on the effective plastic curvature related to the first-yield curvature,
and taking into account the increased strength. “Thus:Chapter 4. Analysis Tools for Direct Displacement-Based Design 151
Cracking: A, =$,H?/3, F,=M.,/H (4.35)
First Yield: A',=6',(H+Lsp) (3; F,=M,/H (4.36)
After Yield: aaa, Mal gg, L,H; F=M/H 437
M, \ M,
Again the accuracy can be improved by replacing Hin Eq,(4.37) by the distance from
the centre of the plastic hinge length to the point of contraflexure,
In both bilinear and refined representations, the displacement ductility capacity is
related to the nominal yield displacement (Eq,(4.33):
Hy=A,/A, 438)
4.2.9 Computer Program for Moment-Curvature and Force-Displacement
The CD provided with this book includes a computer program, CUMBIAM™), for
‘moment-cervature and force-displacement analysis of reinforced concrete members of
circular or square section, based on the principles and equations outlined in the previous
sections. A manual for operating the program is also provided,
4.3 FORCE-DISPLACEMENT RESPONSE OF STEEL MEMBERS
As noted earlier, it will rarely be necessary to carry out moment-curvature analyses of
ceel sections, since the elastic stiffness is unaffected by cracking, and the yield moment
and plastic moment can be readily calculated, However, it should be emphasised that in
calculating the plastic moment capacity, strain-hardening should be considered, and the
‘evel of maximum useable compression strain should be restricted to about 0.02 to avoid
‘ocal buckling of flanges etc. The principles developed above for moment-curvature and
force-displacement analysis of concrete sections can of course be equally applied to steel
sections. This may be advisable when unsymmetrical sections are considered,
4.4. ELASTIC STIFFNESS OF CRACKED CONCRETE SECTIONS
Ir was mentioned in Chapter 1 that, contrary to common assumptions made in force-
based seismic design, the elastic stiffness of cracked concrete sections is essentially
proportional to strength, and the concept of a constant yield curvature independent of
serength is both valid, and important in terms of direct displacement-base design. A
summary of the research leading to these statements, and to Fig.1.4 is included below. A
more complete presentation is available in [P3]. The research is based on moment-
curvature analysis of different concrete sections, and the bilinear representation described
1n Section 4.2.7. It has been verified in numerous experiments!152, Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
4.4.1 Circular Concrete Columns
Circular reinforced concrete columns are the most common lateral force-resisting
elements for bridges in seismic regions, In order to investigate the effective stiffness
of circular columns, a parameter analysis was carried out varying the axial load ratio and
flexural reinforcement ratio for a typical bridge column, ‘The following basic data were
assumed:
© Column diameter D = 2m (78.7 in)
* Cover to flexural reinforcement = 50: mm (2in)
* Concrete compression strength f'ze = 35 MPa (5.08 ksi)
© Flexural reinforcement diameter dy = 40 mm_ (1.575 in)
* Transverse reinforcement: spirals = 20mm (0.79in) ac 100mm (din) spacing
# Steel yield strength Sre = 450 MPa (65.3 ksi
* Axial load ratio Neorg = 0100.4 (9 levels)
© Flexural reinforcement Ratio py/Ag = 0.005 to 0.04 (5 levels)
A selection of the moment-curvature curves resulting from analysis with the program
CUMBIA provided on the attached CD is shown in Fig.4.10 for two levels of flexural
reinforcement ratio, and a range of axial load ratios. Only the initial part of the moment-
curvature curves has been included, to enable the region up to, and immediately after
yield to be clearly differentiated. Also shown in Fig. 4.10 are the calculated bilinear
approximations for each of the curves. Note that the apparent over-estimation by the
bilinear representations of the actual curves is a function of the restricted range of
curvature plotted, and is resolved when the full curve is plotted. It will be seen that the
moment capacity is strongly influence by the axial load ratio, and also by the amount of
reinforcement. However, the yield curvature of the equivalent bilinear representation of
the moment-curvature curves does not appear to vary much between the curves.
Data from the full set of analyses for nominal moment capacity, and equivalent bilinear
yield curvature are plotted in dimensionless form in Fig.4.11. The dimensionless nominal
moment capacity and dimensionless yield curvature are respectively defined as
My
ov Dp (4.39)
and .
py = 9,DI€, (4.40)
where & = fylEs is the flexural reinforcing steet yield strain
‘The influence of both axial load ratio and reinforcement ratio on the nominal moment
capacity is, as expected, substantial in Fig. 4.11(2), with an cight-fold range berween
maximum and minimum values. On the other hand, it is scen that the dimensionless yield
curvature is comparatively insensitive to variations in axial load or reinforcement ratio.Chapter 4. Analysis Tools for Direct Displacement-Based Design
153
Dimensionless Moment (My/(eD!)
40000
10000
0 0.002 0.004
Curvature (1/m)
0.006
(a) Reinforcement Ratio = 1%
50000
40000
30000
Moment (kNm)
8
8
40000
0 0.002 0.004
Curvature (I/m)
0.006
(b) Reinforcement Ratio = 3%
Fig.4.10 Selected Moment-Curvature Curves for Circular Bridge Columns
(D = 2m; Pee = 35 MP2; fe =450 MPa)|P3)
0.6
0.2
0.08
0.04
0 of 02
03
Axial Load Ratio (N,/f'eA,)
4
(2) Nominal Moment
Ave. H0%
Average 0,D/e,= 2.25
0s
Dimensionless Curvature (9,D/e,)
0
o1 02 03
Axial Load Ratio (N,/FA,)
os
(b) Yield Curvature
Fig.4.11 Dimensionless Nominal Moment and Yield Curvature for
Circular Bridge Columnst51154 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Thus the yield curvature is insensitive to the moment capacity. The average value of
dimensionless curvature of @py = 2.25 is plotted on Fig 4.11(b), together with lines at
10% above and 10% below the average. It is seen that all data except those for low
reinforcement ratio coupled with very high axial load ratio fall within the +10% limits.
It should be noted that though the data were generated from a specific column size
and material strengths, the dimensionless results can be expected to apply, with only
insignificant errors, to other column sizes and material strengths within the normal range
expected for standard design. The results would not, however apply to very high material
strengths (say f->SOMPa (7,25ksi), or f;>600MPa (87ksi)) due to variations in stress-
steain characteristics,
‘The data in Figs. 4.10 and 4.11 can be used to determine the effective stiffness of the
columns as a function of axial load ratio and reinforcement ratio, using Eq, (4,26). The
ratio of effective stiffness to initial uncracked section stiffness is thus given by
Elg _
Elin, GEL,
‘goss
(4.41)
Results are shown in Fig. 4.12 for the ranges of axial load and reinforcement ratio
considered. It will be scen that the effective clastic stiffness ratio varies between 0.13
and 0.91. For the most common values of the variables, however, the ratio will be
between 0.3 and 0.7.
g & ®
L L L
Stiffness Ratio (EI/Elgross)
°F
0 out 02 93 o4
Axial Load Ratio (Ny/f"cAg)
Fig.4.12 Effective Stiffness Ratio for Circular Bridge Coiumns|?3Chapter, Analysis Tools for Direct Displacement-Based Design 155
It should be noted thar for convenience in computing the stiffness ratios of Fig 4.12,
the gross stiffness of the uncracked section has been calculated without including the
stiffening effect of the flexural reinforcing steel. That is
aD
T prose = (4.42)
sos Gy 4.42)
Since the reinforcement increases the uncracked section moment of inertia by as much
as 60% for the maximum steel ratio of 4%, the stiffness ratios related to true un-cracked
sections would be lower, particularly for the higher reinforcement ratios. ‘The value of
the concrete modulus of elasticity used in computing Fig, 4.12 was
S000) 7". (MPa) ; E
E
606007". (psi) (4.43)
4.4.2. Rectangular Concrete Columns
Ductile rectangular columns can occur in bridge design, and at the base level of multi-
storey frame buildings. For the purposes of this study the special case of a square column
with flexural reinforcement evenly distributed around the perimeser was investigated.
The following basic data were assumed:
© Column dimensions b=h=16m (635 in)
© Cover to flexural reinforcement mm (2in)
© Concrete compression strength free = 35 MPa (5.08 ksi)
Flexural reinforcement diameter 2 mm (1.26 in)
© Transverse reinforcement: hoops 20mm dia. (0.79in) /5 legs per layer
© Steel yield strength Je = 450 MPa (65.3 ksi)
© Axial load ratio NAP Ag = 010 0.4 (9 levels)
© Flexural reinforcement ratio py/Ag = 0.005 to 0.04 5 levels)
Moment-curvature trends predicted by CUMBIA for the rectangular sections followed
the same trends apparent for circular columns?
Data from the full set of analyses for nominal moment capacity, and equivalent
bilinear yield curvature are plotted in dimensionless form in Fig.4.13, ‘The dimensionless
nominal moment capacity, and dimensionless yield curvature are respectively defined as:
(4.44)
and
(4.45)56 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Ave. +10%
Dimensionless Curvature (9,h/€,)
Dimensionless Moment (My/{".bh?)
1 Average o)h/e, = 210
oa ss
05
oF T T T 1 oF T T T 1
0 01 02 03 04 0. of 02 03 ot
Axial Load Ratio (N,/f'A,) Axial Load Ratio (N,/f',A,)
(a) Nominal Moment (b) Yield Curvature
Fig. 4.13 Dimensionless Nominal Moment and Yield Curvature for
Large Rectangular Columns!"!
where b and ft are the column width and depth respectively.
‘Trends for the rectangular columns, apparent in Fig.4.13, are similar to those displayed
in Fig.4.11 for circular columns. Nominal moment capacity is strongly dependent on both
axial load ratio and reinforcement ratio, with approximately an eight-fold increase in
moment capacity from minimum axial load and reinforcement ratio to maximum axial
load and reinforcement ratio. Dimensionless yield curvature is only weakly dependent on
axial load ratio and reinforcement ratio, thus implying that the yield curvature is
insensitive to the nominal moment capacity. The average value of dimensionless
curvature of py =2.10 is plotted on Fig.4.13(b), together with lines at 10% above and
10% below the average. It is seen that all data except those for P; =0,005 at both low
and high axial load ratio fall within che +/- 10% limits of the average value.
As with the circular columns, the dimensionless results of Fig.4.13 can be expected to
apply to other column sizes and material strengths within the normal range of material
strengths. Small errors can be expected for small column dimensions, where the ratio of
cover to core dimensions will be significantly larger than for the data presented here. As
with the circular column data, results should not be applied to rectangular columns with
very high strength concrete or reinforcing steel.
‘The data of Figs.4.13 have been used to develop curves for the effective section
stiffness ratio, based on Eq.(4.41). Results are presented in Fig.4.14, For ease of
application of the resulks, che stiffness of the gross uncracked section was computed
ignoring the stiffening cffect of flexural reinforcement as, Igross = bh'/12, with the
modulus of elasticity given by Eq.(4.43).Chapter 4. Analysis Tools for Direct Displacement-Based Design 187
Stiffness Ratio (EI/Elgross)
04
04 02 03
Axial Load Ratio (Nu/f'cAg)
Fig.4.14 Effective Stiffness Ratio for Large Rectangular Columns|51
“The range of effective stiffness calculated in accordance with Eq,(4.41) is from 0.12 t0
9.86 times the gross section stiffness, indicating the strong dependence of effective
stiffness on axial load ratio and reinforcement ratio. Clearly the common assumption of a
constant section stiffness independent of flexural strength is entirely inappropriate.
Results from Fig.4.14 can be applied to other column sizes than those used to generate
the graph by appropriate substitution of section dimensions into Eq,(4.41)
4.4.3 Walls
(®) Rectangular Concrete Walls: Similar calculations to those reported in the
previous two sections can also be carried out for rectangular structural walls, and have
been fully presented elsewhere for both concrete and masonry walls(’2"47l. Analyses for
concrete walls considered two separate cases — one where the flexural reinforcement was
distributed uniformly along the wall length, and the second, more common case where
most of the flexural reinforcement was concentrated at the wall ends, with a
comparatively light amount of reinforcement distributed along the wall length. It is
emphasized that though the latter is the more common case, this is largely because of the
misconception that concentrating the flexural reinforcement at the wall ends increases the
flexural capacicy when compared with the same total amount of reinforcement distributed
uniformly along the wall length. In fact, the flexural strength associated with the two
distributions will be very similar, as simple trial calculations will show.158 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Results are presented in Fig-4.15 for the dimensionless yield curvature for walls with
different axial load ratios and reinforcement ratios, in similar fashion to the above
analyses for columns. However, lower ranges of axial load ratio (with a maximum of
N,/PeoAg= 0.125), and reinforcement ratio (with a maximum of pj = 0.02) were adopted
since these were considered to be practical upper limits for structural walls. Two different
conditions were considered, In the first, a reinforcement ratio of 0,005 was considered to
be uniformly distributed along the wall length, with the remainder concentrated near the
wall ends. In the second case the full amount of reinforcement was uniformly distributed.
In Fig.4.15 the yield curvature has been made dimensionless by multiplying by the wall
length 4, and dividing by the yield strain & of the flexural reinforcement, in similar
fashion to columns
“Average values of dimensioniess curvature of Op,
distributed rebar are plotted in Fi
85 and 2.15 for concentrated and
15 (a) and (b) respectively. An average value of
Poy = Ol (E, =2.0 15% (4.46)
essentially covers all data from the concentrated and distributed analyses. We recommend
that this is sufficiently accurate to be used for design of all rectangular walls.
px=0.005
22 Pio 22
0018
>| 9020
2 2
Dimensionless Curvature (@yl4/E,)
Dimensionless Curvature (ylu/ey)
hea
18 * 18
16 r 1 1 1.6 +f ——— T
0 0.04 0.08 0.2 o 0.04 0.08, a2
Axial Load Ratio (N/f'eAg) Axial Load Ratio (N/f'eAg)
(a) Rebar Concenteated at Wall Ends (b) Rebar Distributed along Wall
Fig.4.15 Dimensionless Yield Curvature for Rectangular Walls
Analyses where all the flexural reinforcement was distributed uniformly along the wall
length resulted in an average dimensionless curvature approximately 10% higher than
given by Eg,(4.46), with about twice the scatter of Fig.4.15IP2", The effective stiffness for
rectangular walls can thus be calculated, as a fraction of gross wall stiffness asChapter 4. Analysis Tools for Direct Displacement-Based Design 159
M, 1
EL El gas = v (4.47)
BS bey (Uy) EUS 12, aa
where ¢is the wall width, and My is the calculated nominal flexural strength. ‘Typical
values are about 0.2 to 0.3. In Eq, (4.47), py may be taken from Eq, (4.46) for all wails
or the more accurate averages from the analyses of the walls with concentrated or
distributed reinforcement may be adopted.
(b) Rectangular Masonry Walls: Similar analyses to those described above have been
carried out for masonry walls!*"l The coefficient for the dimensionless curvature defined
by Eq.(4.46) was found to vary between 2.06 for unconfined concrete masonry with
uniformly distributed flexural reinforcement to 2.17 for confined clay or concrete
masonry, also with distributed flexural reinforcement. It is recommended that an average
value of 2.10 be used in all cases.
(c) Flanged Walls: Paulay?® has investigated the dimensionless yield curvature for
different sections, including flanged concrete structural walls. For I-section and T-section
walls with the flange in compression, he recommends dimensionless curvature
coefficients (Eq.(4.46)) of 1.4. Paulay’s values for rectangular walls are on average about
10% below the values established above. The main reason for this difference appears to
be some simplifying assumptions made by Paulay, including ignoring the effects of strain-
hardening of the flexural reinforcement, which is included in our analyses. Spot
comparisons for flanged walls indicate a similar influence of strain-hardening, and we
therefore recommend a coefficient of 1.5 for flanged walls where the flange is in
compression. Note that for T-section walls with the flange in tension, or I-section walls
loaded perpendicular to the web, the higher values defined in Section 4.4.3(a) above are
recommended.
4.4.4 Flanged Reinforced Concrete Beams
Similar studies have been carried out to investigate the influence of flexural
reinforcement ratio on the stiffness of flanged beams. These were based on the section
shown in Fig. 4.16, and have been fully reported elsewhere"231, The appropriate value
for the stiffness of a beam in a building frame under seismic action will be the average of
the values applicable for positive and negative bending, as a result of the moment reversal
along the beam length. As with columns and walls, it was found that the flexural strength
and stiffness increased essentially proportionately as the reinforcement content increased,
and that effective stiffness ratios (related to the gross section stiffness) varied between
0.17 and 0.44.
If strain-hardening is ignored, and based on the average of positive and negative
bending, it was found that the dimensionless curvature could be written as160 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
1560
4 legged R10
@ dors
Fig.4.16 Reinforced Concrete Beam Section Analysed for Yield Curvature!)
bo, = 0h, /€, =1.7£10% (4.482)
Including the effects of strain-hardening incteased the dimensionless yield curvature to
Op, = O,hy /€, = 1.910% (4.480)
For rectangular-section beams rather than flanged beams, average values for
dimensionless curvature were about 10% higher than given by Bq (4.48), and with
somewhat increased scatter. It is thus clear that the concept of a constant dimensionless
yield curvature fox concrete beam sections is an adequate approximation.
4.4.5. Steel Beam and Column Sections
‘There is less need for studies on steel sections, of the type teported in the previous
sections for concrete and masonry members, since the elastic section stiffness can be
directly computed from the known section dimensions. However, at the start of the
design process, the overall depth of the section may be approximately known, but the
flange thickness, and hence the strength and stiffness of the section will be chosen after
strength requirements have been computed. In force-based design, a trial and error
approximation should be carried out to ensure assumed stiffness and final flexural
strength are compatible. For direct displacement-based design, however, the initial yield
curvature is of greater interest, as it defines the yield displacement, and hence the ductility
demand, enabling the effective damping to be computed
It will be readily appreciated that for a symmetrical structural steel section, of overall
depth hr the first-yield curvature will be given by:Chapter 4, Analysis Tools for Direct Displacement-Based Design 161
(4.49)
This curvature applies at the first-yield moment. For typical steel beam and column
sections the nominal flexural strength will be only slightly higher than the yield moment
say My = 1.1 Mj), and hence the effective yield curvature can be written as:
&y
6, =2.25% (4.50)
Icis of interest to note that this value is almost identical to concrete column sections
reinforced with steel of the same yield strength.
44.6 Storey Yield Drift of Frames
‘The comparative invariance of dimensionless yield curvature of beams and columns
indicates that storey yield drift of frames might similarly be essentially independent of
reinforcement ratio and strength. Fig.4.17(a) shows a typical concrete beam/column
subassemblage extending half a bay width either side of the joint, and half a storey height
above and below the joint. ‘This can be considered a characteristic element of a frame
building, Since bay width will normally exceed storey height, and column curvatures will
wypically be less than beam curvatures as a consequence of capacity design procedures
see Section 5.8), beam flexibility is likely to be the major contributor to the deformation.
‘The deflected shape is shown in Fig.4.17(b). The yield drift 8,can be expressed as
6, =8,, +0, +2A,/L, +2A,/L, (451)
where Oy and G are the rotations of the joint centre due to beam flexure and joint shear
deformation respectively, Ac is the flexural deformation of the column top relative to the
cangent rotation at the joint centre, and A, is the additional deformation of the column
cop due to shear deformation of beams and columns. To allow for strain penetration of
longitudinal reinforcement into the joint region, it is assumed that the yield curvature in
the beam develops at the joint centroid, and reduces linearly to zero at the beam midspan,
as shown in Fig.4.17(¢).
The yield drift due to beam flexure is thus:
G(0-5L,) _ Oly
* 3 6
For a concrete frame, ignoring strain-hardening, and thus substituting from
Bq (4.48(@)):
6, (4.52)162 Priestley; Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
8, =0.283e, [2] (4.53)
ly
‘Typical calculations based on 2 storey height/bay length ratio of 0.533 (storey height
= 3.2m, bay length =6m) and a maximum column curvature of 0.754, indicate column
displacement de will add about 40% to the yield drift in Eq, (4.51). Itis further assumed,
based on experience, that the joint deformation and member shear deformation add 25%
and 10% respectively to the yield drift. As a consequence, the yield drift for a reinforced
concrete frame may be estimated as
8, =(1.0+0.4+0.25 +0.1)x name, |
5
L,
.5é,| —* 4.54)
2] (4.54)
Equation (4.54) is compared in Fig4.18 with the results from 46 beam/column test
assemblages which included a wide range of possibly relevant parameters”, including
* Column height/beam length aspect ratio (H_/Ls) : 0.4 ~ 0.86
© Concrete compression strength (Pr) 122.5 — 88MPa (3.3 — 12.8ksi)
* Beam reinforcing steel yield strength (fre) 276 ~ 611MPa (40 — 89 ksi
© Maximum beam reinforcement ratio (A’/bu)—: 0.53% —
© Column axial load ratio (NAP ey) 20 - 0.483
© Beam aspect ratio (Li/ly) 44-126
Test units with equal, and with unequal top and bottom reinforcement ratios were
considered, as were units with and without slabs and/or transverse beams framing into
the joint. Note that Eq.(4.54) only includes two of these parameters (beam reinforcement
yield strength (which dictates the yield strain & =/,/E,), and beam aspect ratio), on the
assumption that the other parameters are not significant variables. As is apparent from
Fig.4.18, the agreement between experimental drifts and predictions of Eq.(4.54) is
reasonable over the full, and rather wide range of yield drifts. The average ratio of
experiment to theory is 1.03 with a standard deviation of 0.16, Considering the wide
range of parameters considered, the comparatively narrow scatter is rather satisfactory.
Note that the wo test units with experimental drifts of about 1.5% are thought to have
suffered beam slip through the column joint region, resulting in excessive yield drift, but
have been included in the averaging, The significance of different experimental
parameters included in the list of variables noted above, to the theoretical/experimental
drift ratio has been examined elsewhe-ei®."™, In no case was the significance of any of
the parameters not included in Eq.(4.54) found to be high enough to warrant inclusion in
the design equation.Chapter 4. Analysis Tools for Direct Displacement-Based Design 163
(c) Assumed Beam Curvature Distribution
Fig. 4.17 Elastic Deformation Components to Drift of a
Beam Column Subassemblage!™164 Priestley, Calvi and Kowalsky.
isplacement-Based Seismic Design of Structures
Experimental Drift Ratio (%)
0 05 1 15
Theoretical Drift Ratio (%)
Fig. 4.18 Experimental Yield Drifts of Reinforced Concrete Beam/Column
Test Units Compared with Predictions of Eq.(4.54)("23
cis evident that the procedure adopted above to estimate the yield drift of reinforced
concrete frames can also be adapted to predict the yield drift of structural steel frames.
Substituting from Eq.(4.50) into Eq.(4.52), the drift due to beam flexure will bet
6, = once) | (4.55)
5
and making the same assumptions about the percentage increase of yield drift from joint
rotation, column flexure, and shear deformation, the following drift predictor results
L 5,
O, =0.65e, (4.56)
hy
That is, the yield drift for a structural steel frame is expected to about 30% higher than
for a reinforced concrete frame with the same gross dimensions, and the same steel yield
Stress. This equation has not, however, been checked against experimental results.
4.4.7 Summary of Yield Deformations.
From the moment-curvature analyses reported in the previous sections, it was found
that stiffness and strength are effectively proportional, for a given structural member type
and size, or structure type. The independent parameter, for stiffness caleulations, is thusChapter 4. Analysis Tools for Direct Displacement-Based Design 165
yield curvature, or yield displacement. The following yield curvatures are applicable
for the “corner” of the equivalent bilinear approximation of force/ deformation response:
© Circular column: 6, =2.25e,/D (4.578)
Rectangular column: 9, = 2.10, /h, (4.57)
* Rectangular concrete walls: 9, = 2.00E, /1, (4.570)
* = T-Section Beams: @, =1.70e, /h, (4.574)
© Flanged concrete walls: @, =1.50e, /1, (4.57e)
© Rectangular masonry walls: @, =2.10E, /I, (4.578,
The equation for flanged walls applies for I-section walls and for T-section walls when
the flange is in compression. For reinforced concrete and structural steel frames, the yield
drift can be expressed, with adequate accuracy as
L,
Concrete frames: 8, =0.5€,— (4.58a)
ts
Structural steel frames:
y
0.65e, [ ~. ] (4.58b)
45 ANALYSES RELATED TO CAPACITY DESIGN REQUIREMENTS |.
In Section 3.9, capacity design was introduced. The basic philosophy is that DDBD is
used as a means to determine the required strength of locations where inelastic rotations
‘plastic hinges) are intended. To ensure that plastic hinges do not occur at other parts of
the structure, and to ensure that undesirable modes of inclastic deformation, such as
shear failure do not develop, the dependable strength of these locations and actions is set
to be higher than the force levels at these locations, corresponding 10 the maximum
feasible strength being developed at the plastic hinges. This is in recognition that in
ductile design, it is the actual strength, rather than the conservative design strength, that
will be developed in the design-level earthquake.
This premise is illustrated in Fig. 4.19, which examines the force-displacement
response of a simple bridge pier under lateral seismic response. If the pier had very high
strength, it could respond clastically to the inertia forces, and have the maximum
displacement and force corresponding to point A in Fig.4.19(c). The design strength is
however much lower, and if the actual strength exactly equalled the required strength, the
expected maximum response is defined by point B. If the actual stsength exceeds the
design strength, thea it will be this strength that is developed, unless the actual strength166 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
exceeds that corresponding to point A. The cesponse corresponding to actual strength is
defined by point C. Note chat the different strengths of A and B imply different stiffness.
F La}
— >
(a) Structure _(b) Deflection (c) Force-Displacement
Profile Response
Fig.4.19 Strength Developed in Design Earthquake by a Simple SDOF Pier
‘There are a number of reasons why the actual flexural strength may exceed the design
strength:
* Material strengths (c.g. concrete compression strength, steel yield strength) may
exceed the nominal or characteristic values used in design.
* Dependable flexural strength may incorporate a strength reduction factor (or
partial material factors)
* Suainchardening of reinforcement or structural steel may not have been
considered in determining the flexural capacity of the section.
* The section size or reinforcement content may exceed the exact values required
to equal the required strength.
A section or action being capacity-protected would need to take these possible
increases of the plastic hinge flexural strength into account, and be designed for che
appropriate action in equilibrium with the enhanced strength of the plastic hinge, In
addition to this, the basic design for the plastic hinges may be based on a SDOF estimate
of response, as is the case with DDBD. Amplification of che action requiring capacity
protection due to higher mode effects must also be taken into account.
‘The general requirement for capacity protection is defined by Eq.(3.61), which is
reproduced here as E.q.(4.59) for convenience:
OsSp 2Sp =P OS, (459)
where § is the value of the design action being capacity protected, corresponding to the
design lateral force distribution found from the DDBD process, ff is the ratio of
overstrength moment capecity to required capacity of the plastic hinges, @ is theChapter 4. Analysis Tools for Direct Displacement-Based Design 169
Design Verification and Capacity Design Factors: \ moment-curvature analysis
is now run with the design longitudinal reinforcement of 52D40 bars (65,300mm?), and
the transverse hoops of D20@120 crs. using the expected material strengths. The results
are shown in Fig4.20 by the line identified as “Design”. At the design curvature of
11,0274/m the moment capacity is Mp = 27,200kNm (241,000 kip.in). This is 1.2%
above the required design value.
Also shown in Fig.4.20 are qvo moment-curvature curves designated “Overstrength”
A and B. Curve A corresponds to maximum feasible overstrength. For this case, fo
=1.7f. = SIMPa (7.4ksi); fp = 1.3; = 546MPa (79.1ksi); and fyio = 1.3fon = 546MPa
~9.1ksi), in accordance with the recommendations of Section 4.2.4. At the design
curvature of 0.0274, the moment capacity is 31,900kNm (283,000 kip.in), which is
18.7% higher than the design requirement. Thus the appropriate value of @ to use in Eq,
4.59) is @ = 1.187, This would be appropriate if it were used to determine required
strength in other members of the bridge (say the superstructure). It would, however, be
inconsistent to use this value of overstrength for determining the required shear strength
of the column, since it is based on maximum feasible strengths of concrete and transverse
xeinforcement, whereas the shear capacity will use lower values. It is clear that the same
values should be used for concrete strength and transverse reinforcement yield strength
for both overstrength demand, and shear capacity since they apply to the same region of
the structure.
30000
20000
Moment (kNm)
10000
° 0.01 0.02 0.03
Curvature (m=!)
Fig.4.20 Moment-Curvature Response for Example 4.1170 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Accordingly, Fig, 4.20 shows a second overstrength curve (B), which is based on Pe»
See = 39MPa, and fyio = fon =420 MPa, since these values should be used in estimating
the shear capacity (see Section 4,7). It will be seen that this results in a rather small
reduction in the overstrength moment capacity, and little inefficiency would result from
using curve A. At the design curvature, the moment capacity of curve B is 30,840kNm
(272,900 kip.in), which implies g = 1.146.
An additional curve, labelled “(Force-Based)” is included in Fig, 4.20. This is based on
conventional force-based design assumptions that the minimum specified material
strengths (P'o fp fys) should be used in design, and that strain-hardening of the flexural
reinforcement should be ignored. The nominal capacity would be estimated at an extreme
fibre compression strain of 0.003. For the material properties and reinforcement details
used in the final design above, this would result in a nominal flexural strength of My =
20,400kNm (180,500 kip.in), which is well below the required design strength. Using
the same assumptions about material strengths, it would be necessary to increase the
longitudinal reinforcement to 93,650 mm? (a 48% increase) to match the required
strength. The resulting overstrength demand can be estimated by the ratio of the capacity
of curve A to the nominal strength of the (Force-Based) curve, resulting in
& = 31,943/20,800 = 1.536.
Ifa flexural strength reduction factor of gy = 0.9 were incorporated in the design, as is
common practice in many codes, the overstrength factor would increase to @ = 1.706.
‘The extreme and unnecessary conservatism of current force-based design procedures is
thus apparent in this example,
4.5.2 Default Overstrength Factors
For some simple structures the design effort involved in determining the required
flexural overstrength factors for capacity design may be excessive. In these cases it would
be permissible to use conservative default values for the overstrength factor @*, Provided
the design is based on a strain-hardening model for the flexural reinforcement, it is
recommended that the default value should be g = 1.25. If strainchardening is ignored
in determining required section properties, it is recommended that g = 1.6 be assumed
4.5.3, Dynamic Amplification (Higher Mode Effects)
Dynamic Amplification due to higher mode response is dependent on the type of
structure being designed. As such, it is separately dealt with in each of the special chapters
devoted to different structural types. Note that in Ex. 4.1 there were no higher modes,
4.6 EQUILIBRIUM CONSIDERATIONS IN CAPACITY DESIGN
Equilibrium must be satisfied at all stages of design, whether design level or
overstrength is considered. This is illustrated in Fig4.21 for a simple portal frameChaprer 4, Analysis Tools for Direct Displacement-Based Design 171
subjected to both gravity and seismic loading, at design and overstrength levels. It is
ssumed that the columns are ductile, and reach their capacity in both design and
iverstrength response, as discussed in Section 4.5. For simplicity the moment capacities
cop and bottom of the columns are assumed to be identical
(a) Design Moments and Forces (b) Overstrength Moments and Forces
Fig.4.21 Equilibrium of Moments and Forces in a Portal Frame
(moments drawn relative to member centrelines on tension side of member)
(a) Design-Level Response: The design lateral force Fp is in equilibrium with the sum
of the shears in the two columns. ‘The design moments in the columns form at the
column base and at the Soffit of the beam, Since the shear in the columns is the slope of
the bending moment, the following equilibrium equation applies:
2My, +2M yy
A
Fy (4.61)
where His the clear height to the soffit of the beam, and Mp, and Mpp are the design
moment capacities in the left and right columns respectively. Note that Mp, # Mor, since
the axial forces in the two columns are different. Note also that gravity moments have no
influence on the validity of Eq,(4.61), since the column capacity is developed regardless
of the ratio of seismic and gravity moments in the columns.
To determine the moment capacities of the two columns, it is thus necessary to
determine the axial forces in the columns.
Vertical force equilibrium requires that the axial forces in the columns must equal the
beam shears at the keam/column joint centroids. To find the beam shears we first note
that the beam and column moments must also be in equilibrium at the joint centroids,12 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
and that the column moments at the joint centroids exceed the moments at the beam
soffit
(0.5H +0.5h,)
OSH
My = Mp (H +4,)( (4.62b)
My, =My, = Mp (H+h,)/H (4.62a)
Note that the gravity load moments resulting from the load Wagain have no influence
on the moments at the joint, which are torally defined by the column moment capacities,
‘The shear in the beam can be separated into the gravity load and seismic force
‘components:
Gravity: V,=W/2 (4.632)
(My, + Mop)
L
‘The sign of the gravity-load shear changes at the evo ends of the beam, while the
seismic shear is constant, ‘The reactions at the base of the columns (ignoring the column
weights) are thus:
a Mn My
2 L
Seismic: V, = (4.63b)
Rog = (4.64)
W (My + May)
2 L
The moment capacities of the column hinges must be calculated with the appropriate
value of the axial force, given by Fq,(4.64). This may involve iteration, though normally
one or two cycles arc sufficient. In the initial design phase, when the design force Fp is
known, and the column design moments are required, it is useful to recall that the
moment-axial force relationship is likely to be nearly linear over the range of axial force
expected, and hence the required flexural reinforcement content (for a reinforced
concrete design), can be found from the approximate combination:
N=R,. =W/2 (4.65)
This should then be checked with a detailed analysis, using the selationships of Eq.(4.62)
and (4.64).
(b) Overstrength Response: The principles utilised in the previous section indicate
that when the overstrength moment capacity is estimated at plastic hinge locations, the
axial force should be adjusted to reflect the increased beam seismic shear resulting from
the increased plastic hinge moment capacity of the columns. With respect to Fig. 4.21(b),
the following relationships apply for seismic response force and seismic beam shear
respectively:Chapter 4. Analysis Tools for Direct Displacement-Based Design 173
2(m?+M2)
H
o (M2 +Me)-(1+h,/H)
a
The seismic overstrength moment capacities, and the overstrength column shear
demands for capacity design thus need to be determined in conjunction with the revised
column axial forces:
Fe= (4.66)
(4.67)
RO =W/2-VP; Ro=W/24+V2 (4.68)
Required shear strength for the columns is given directly from the overstrength
moment distribution as:
V2 =2M?P/H; Ve =2Me/H (4.69)
and the beam maximum required shear strength, at the right end of the beam is
Vg=Ve+Wws2 (4.70)
‘The principles applied in this section are simple fundamental requirements of
equilibrium, and should be self-evident to all designers. Unfortunately we find that
equilibrium is not afforded the same emphasis as (Say) matrix analysis methods in
engineering curricula, and have hence provided this rather basic treatment of the subject.
4.7 DEPENDABLE STRENGTH OF CAPACITY PROTECTED ACTIONS
‘The discussion on capacity protection has thus far concentrated on demand. Equation
(4.59) requires that the dependable capacity of the capacity protected action, @sSp is at
least equal to the demand. Note that in this case the use of a strength reduction factor, ds
is justified, since the consequence of the strength of the required action being less than
chat of che demand may be catastrophic. An example is the shear strength of a building or
bridge column, Insufficient shear strength to cope with shears associated with flexural
overstrength in plastic hinges could result in shear failure followed by reduction in
capacity to support gravity loads to the extent that failure occurs.
4.7.1. Flexural Strength
Members where plastic hinges are not prescribed must have adequate flexural strength
to ensure that unexpected plastic hinges do not occur. ‘This is to ensure that undesirable174 Pricstley, Calvi and Kowalsky. Displacement-Based Seismic Design of Sinsctares
deformation modes such as soft-storey mechanisms in building frames do not develop.
Aspects related to this are fully covered in (P1] and Section 5.2.3,
Determination of nominal flexural strength My is straightforward, and basic to all
aspects of structural design, and again will not be further elaborated. However, we
recommend that advanced section analyses, based on moment-curvature analyses, and
including the effects of reinforcement strain-hardening be included in the design.
Capacity should be based on nominal, or characteristic material strengths, rather than the
enhanced values used for determining strength of plastic hinges, defined in Section 4.2.5.
For concrete sections, nominal capacity should be determined at an extreme fibre
compression strain of 0.004, or a reinforcement strain of 0.015, whichever occurs first. A
flexural strength reduction factor of g = 0.9 should he adequate to cope with possible
material understrength, though the value will depend on local code requirements for
material strengths, and expected construction quality.
4.7.2, Beam/Column Joint Shear Strength
Integrity of beam/column joints is essential to the successful performance of building,
bridge, wharf and industrial frames, and requires careful consideration of the force
transfer through the joint region. This has been extensively covered for buildings in (PI)
and for bridges in [P4]. Since an adequate coverage of this important topic requires
considerable length, it is not repeated here, and the interested reader is referred to the
above texts.
4.7.3. Shear Strength of Concrete Members: Modified UCSD Model
‘There is also a great body of information relating to the shear strength of concrete
members in the research literature (e.g. P4, K4, AD etc}, with a significant divergence
becween design approaches and design equations required by different national codes.
Many of these do not recognise that flexural ductility affects the shear strength in plastic
hinge regions, and the influence of axial force resulting from gravity and from prestress is
often treated quite differently. In [P4] a new approach for determining the shear strength
of concrete columns was outlined that considers most of the critical parameters, and
since this has subsequently been upgraded to improve agreement with experimental
resultsik*l, the modified approach, generally referred to as the modified UCSD model is
outlined here. Independent studies have indicated that this approach provides a better
agreement with experimental results than other methods.
In the modified UCSD model, the shear strength of concrete sections is found from
the additive equation:
“es = WV eey =O, We +V 5 + Vp) 4.71)Chapter 4. Analysis Tools for Direct Displacement-Based Design 175
where Vo, Vs, and Vp are shear strengths provided by concrete mechanisms, transverse
reinforcement truss mechanisms, and axial force mechanisms respectively. These are
described separately in the following sections.
(a) Concrete Shear-Resisting Mechanisms, (Vc): The key component of the strength
of concrete shear-resisting mechanisms in flexurally cracked members is provided by
aggregate interlock on the rough flexure/shear cracks. In plastic hinge regions, the
strength of aggregate interlock reduces as the flexure/shear crack widens under ductility
The strength is also dependent on the aspect ratio of the member, defined as the distance
tom the critical section to the point of contraflexure, divided by the section depth, (ie.
M/(VD), where Mand Ware the moment and shear at the critical section and Dis the
oral section depth) and on the volumetric ratio of longitudinal reinforcement, P=Ae/Ag
The strength is thus given by:
KPA = OBS. 084) 472)
where 10sa=3-Me1s (4.72a)
VD
B=0.5+20p, <1.0 (4.72b)
and Yis given by Fig, 4.22, for concrete columns. Note that in Fig.4.22, the prime variable
< the curvature ductility demand, which is directly related to the width of flexure/shear
cracks in the plastic hinge region. .\ secondary variable is the mode of ductility. tt has
deen found!s? that the strength of concrete shear-resisting mechanisms of members
subjected co ductility demands in nwo orthogonal directions (biaxial ductility) degrades
nore rapidly than sections subjected to uniaxial ductility. This is reflected in Fig 4.22.
0.25 03
92
Fo2
01s :
oa g
z Sa Biaxiat
0.05
0 °
o 4 8 on woo 0 4 8 wn
Curvature Doetilty Curvature Ducility
(a) Design of New Members (b) Assessment of Existing Members
Fig.4.22 Ductility Component of Concrete Shear-Resisting Mechanism for
Columns (Modified UCSD Model)176 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Figure 4.22 also distinguishes between design of new members, and assessment of
existing members. A less conservative approach is approptiate for the latter, since the
consequences of excessive conservatism is assessment may be unnecessary strengthening,
For concrete beam sections, it is suggested that the strength given by Eq.(4.72) be
reduced by 20% to compensate for the conditions of concrete confinement which are
less satisfactory than for columns, and that the tension, rather the total reinforcement
area be used to determine B.
Note that the reduction in shear strength with ductility only applies to the plastic
hinge regions of members designed for ductility. In parts of members between plastic
hinges, and in members protected against plastic hinging by a capacity design approach,
the value of ¥ used in Eq,(4.72) will be the value at a curvature ductility of 1.0.
(b) Axial Load Component, Vp: In many design equations for concrete shear strength
the axial load on the section is combined in a composite equation with the concrete
shear-resisting mechanisms. This would imply that the well-known enhancement of shear
strength with increased axial compression would reduce with flexural ductility. This is not
supported by experimenis, In the UCSD model, the shear strength enhancement resulting
from axial compression is considered as an independent component (see H4,(4.71)),
resulting from a diagonal compression strut, as illustrated in Fig. 4.23 for columns, and in
Fig.4.24 for the beam of a portal frame (the extension to other beams with seismic
moment-reversal along the length is obvious). It will be noted that the axial force in
beams will often be low, and it is traditional to ignore its influence on shear strength,
Nevertheless, in some structures, particularly bridge bents, the influence can be
substantial. ‘This will particularly be the case if the beam is prestressed.
P| (Axial Load) P| (Axial Load)
(cracks)
Ve 8 MW
(a) Double Bending (b) Single Bending
Fig. 4.23 Contribution of Axial Force to Column Shear StrengthChapter 4. Analysis Tools for Direct Displacement-Based Design 177
Inertia force
(cracks)
Fr
(prestress)
Fo
(prestress)
Fig. 4.24 Contribution of Axial Force to Shear strength of a Portal Framel?4)
In Figs.4.23 and 4.24, the average inclination of the strut involving the axial force is
shown at an angle £ to the member axis. For the column of Fig. 4.23(a) which is
restrained from rotation at top and bottom, the axial force is effectively applied to the
column through the centre of the flexural compression zone at the beam top, and exits
through the centre of flexural compression at the bottom. The horizontal component of
this strut acts to resist the applied shear force, thus enhancing the column shear strength.
For the cantilever column of Fig4.23(b), the axial Joad is applied at the centre of the
column at the top, but again exits through the centre of Alexusal compression at the
bottom. The angle of the strut to the column axis is less than in Fig. 4.23(a), and hence
the resistance to lateral shear is less. In Fig.4.24, due to the uniformly distributed inertia
force, the axial force varies along the length of the beam, which will be subjected to axial
tension at one end, and axial compression at the other, unless prestressed, as suggested in
Fig4.24, The vertical component of the inclined strut either adds to, or resists the applied
shear force, depending on whether the axial force is tensile or compressive.
In assessment of existing structures the fall axial force component should be relied on.
For new structures @ more conservative approach, where the axial force component is
reduced by 15% has been proposed", ‘The following equations thus apply for Vp
Design: V, =0.85P.tan (4.73a)
Assessment: V, = P.tang (4.73b)
where Fis the angle between the strut and the member axis.
(c) Transverse Reinforcement Truss Shear-Resisting Mechanism. ‘the strength of
transverse reinforcement truss mechanisms is illustrated in Fig.4.25 for rectangular and
circular columns, The rectangular column illustration (Fig.4.25(a)) is also relevant for
rectangular or T-section beams. The critical flexure-shear crack crosses the section at an178 Priestley, Calvi and Kowalsky. Displacement-Hased Seismic Design of Structures
©@o
(9) Rectangular column (b) Cireular columa
Fig.4.25 Effectiveness of Transverse Reinforcement for Shear Resistance of
Columns'?4!
average angle of @ to the vertical axis. The layers of transverse reinforcement crossed by
the crack act to transfer some of the shear force across the crack. The maximum force
that can be transferred by a layer of area Ay depends on the yield strength of the
transverse reinforcement, and the orientation of the bars of the layer with respect to the
axis along which the shear is applied. For example, at the vertical sections denoted 1 in
Fig4.25), the octagonal hoop, which forms part of the layer will provide resistance at
45° to the direction of applied shear, and the effective area of this hoop at this section
will be Ay/V2, where Ay is the area of the hoop bar, whereas the outer peripheral
rectangular hoop will be fully effective at both sections 1 and 2. It can be shown! that
for the case of one peripheral and one octagonal hoop, the average effective area is Ay =
3.614), If however, the transverse reinforcement consists of a peripheral hoop and two
overlapping internal hoops (shown by dashed lines in Fig. 4.25(a)), the average effective
area is Ay = 4.67Ay,
With a circular column (Fig4.25(b)), the orientation of the hoop forces restraining a
flexure shear crack depends on the position of the individual hoop where the crackChapter 4. Analysis Tools for Direct Displacement-Based Design 179
intersects it. At the column centreline, the hoop force Ayfys is parallel to the applied shear
force, but as the distance from the column centreline increases, the inclination arto the
direction of applied force also increases, and the effective force restraining the crack, i
Avfyutan at decreases. Given that there are two forces restraining the crack (Fig.4.25(b)),
the average restraining force, taking all possibie locations of the hoop or spiral with
respect to the intersecting crack, can be shown to be
Aw (474)
If the angle of the flexure-shear crack to the member axis is @ for either the
rectangular or circular column, che depth from the extreme compression fibre x0 the
neutral axis is €, and the cover to the centre of the peripheral hoop is ¢», then the number
of layers of hoop crossed by the inclined crack is
(D-c-c,)
n
cot(A) (4.75)
where sis the spacing of hoops or spirals along the member axis,
The total shear resistance provided by the transverse reinforcement for the rectangular
column can thus be estimated as:
AJfyy(D-¢-¢,)-cot(8)
s
Rectangular column:
(4.76a)
where A, is the effective area of hoops in a single layer, as discussed above, and for a
circular column with circular hoops or spirals:
_t Afy(D=e~€,)-cot(8)
2
Circular column: V
(4.76b)
Equation (4.76a) may be considered “exact”, but Eq.(4.76b) is approximate since the
average effective force given by Eq,(4.74) is exact only if the crack penetrates the full
width of the column, However, it has been shown that the error is small, and
conservativelK4l, Although standard practice in the United States adopts an angle of 8 =
45°, this has been found to be unnecessarily conservative, provided longitudinal
reinforcement is not prematurely terminated. European practice is based on plasticity
theory and a variable angle @. It has been found that sufficiently conservative designs can
be obtained taking 6 = 35°. For assessment of existing structures the less conservative
value of @ = 30° is appropriatel""l, The principles and equations outlined in this section
can also be applied to concrete beam sections.180 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
(d) Strength Reduction Factor for Shear Capacity: For design it is appropriate 10
include a strength reduction factor in addition co the conservatism applied in the
equations for the individual shear-resisting components described above. It is
recommended that this be taken as Js = 0.85 (see Fg.(4.71)). For assessment of existing
structures, where the required shear strength has been determined in accordance with the
capacity-protection measures of Section 4,5, no shear strength reduction factor need be
applied, since a “best estimate” of performance will be required.
(©) Comparison of Assessment Equations with Circular Column Test Data: Figure
4.26 compares shear strength predictions of Eg,(4.71) using the assessment values of the
component strengths, and a streagth reduction factor $s =1.0, with experimental data for
a wide range of circular columns. Brittle and ductile shear failures are those for which che
initial shear strength (curvature ductility =1) was iess than or more than, respectively, the
shear corresponding to flexural strength. Measured, rather than nominal, material
propetties were used in the predictions which are very satisfactory, regardiess of what
variable is used to organize the cest results. Note that the columns failing in flexure have
strengths below the predicted shear strength. This does not indicate unsatisfactory
bebaviour, since the strength was dictated by the flexural capacity which was less than the
shear strength. Design shear strength, including the @s =0,85 factor, is shown by the
dashed line, and is a lower bound to all shear failures.
@ Prediction of Ductility Capacity for Flexure-Shear Failure of Existing Columns:
With new designs, the design procedure should ensure that shear failure does aot occur.
With existing columns, however, ic may be found that shear failure is predicted. When
this occurs at less than the flexural strength of the column, brittle failure is expected,
though the displacements will be larger than assessed from a simple bilinear
approximation based on flexure alonel™, This is explained in relation to Fig.4.27.
Figure 4.27 shows the force-displacement response of bridge columns roughly based
on Example 4.1, The bilinear flexusal response is indicated by the solid line, and the
“refined” response, including onset of cracking and non-lineatity of the post-first yield
behaviour in shown by the dashed line labelled “A”. The second dashed line, labelled “B”
includes shear deformation calculated in accordance with recommendations made in
Section 4.8, Three possible shear strength envelopes are shown, each corresponding to 2
different spacing of the transverse reinforcement. Scrength envelope 1 corresponds to
close spacing of the transverse reinforcement, and shear failure is not expected before the
full displacement capacity is reached, whether shear deformation is included or not.
Shear strength envelope 2 degrades to intersect the predicted force-displacement
response. The failure displacement (and force) depends on whether shear displacement is
included or ignored, The values are 284mm and 297mm (11.2in and [1.Tin) respectively
In terms of displacement-based design the differences are nor significant, particularly
when the slight reduction in strength resulting from the increased displacement associated
with the response including shear deformation is considered. ‘The bilinear and refined
displacement curves are identical in this region.Chapter 4. Analysis Tools for Direct Displacement-Based Design 181
+ Ductile Shear Fares +
uw Brite Shear Failures wa
| 4 Fleer (oo shear aire)
gi . gl :
z oy ty te z
£ te taast 00 tte B a +
2, Les £35054 07, ao £, ait * *
= = oF wae ae =
Foe, Fos & * *
4.
ey
06 06 4 :
To] t T T 1
o 4 8 2 6 2% ° 4 8 n
(@)Test Number (&) Displacement Duct a¢ Failure
324 ge
a é
21 gi
| e
Fos 4 08 “4
ve 4 oo 4 .
1 T 1 1 1 1 1
0 om 002 003m 01 oo 0.03
(©) Longitodinal Reinforcement Ratio. (a Teanovese Reinforcement Ratio
1
L
ge qi
st a
eit 4 +4 ea
Fos 8 ‘ Soa
b
L
01 0 Of 0203 Od 120146 2 ngs
(6) Axial Load Ratio (N/t"AQ) (0) Aspect Ratio (M/VD)
Fig.4.26 Shear Strength Comparison: Circular Column Data vs Assessment
Prediction Organized by Significant Variables (Design Strength Shown Dashed)
With shear strength envelope 3, corresponding to wide-spaced transverse
reinforcement, there is a significant difference in the predicted failure displacements
depending on which force-displacement curve is used, with the values being 100mm,
135mm and 149mm (3.9in, 4.3in, and 4.9in) for the bilinear, refined, and refined+shear182 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
deformation respectively. In displacement based assessment of an existing structure the
higher displacements would translate into a significant increase in seismic intensity able to
be accommodated by the structure, particularly when the enhanced damping associated
with the implied ductility is included in the analysis. Note that shear deformation has
conservatively not been applied to the strength envelopes in the above discussion
= Strength 1
Shear Force (MN)
° 100 200 300 400 500
Displacement (mm)
Fig.4.27 Force-Displacement Response of Columns with Different Shear Strengths
4.7.4 Design Example 4.2: Shear Strength of a Circular Bridge Column,
The bridge column of Design Example 4.1, presented in Section 4.5.1, is analysed 0
see if i¢ has adequate shear strength to satisfy the design requirements of Sec. 4.7.3.
Additional information required is that the column is expected to have similar ductility
demands in orthogonal directions, and hence the biaxial data of Fig.4.22 apply for
determining the strength of the concrete shear-resisting mechanisms
Design shear force: Curve B of Fig.4.20, which is based on overstrength flexural
reinforcement properties, but expected concrete strength, applies. Thus, as noted in
Section 4.4.1, the overstrength moment capacity is 30.8 MNm (273,000 kip.in), With a
colums height of 10m (32.8%), the overstrength shear force is thus:
V" = 3.08MN (693kips).
A shear strength reduction factor af gy = 0,85 is specified. There is no dynamic
amplification to be applied, and hence the capacity-design Eq.(4.71) can be expressed as:
+V,4V,2—
,
v,
63 MN (8 1 Skips)Chaptee 4. Analysis Tools for Direct Displacement-Based D: 183,
Concrete shear-resisting mechanisms: From Ex.4.1, the curvature ductility is:
p/G, = 0.0274/0.00289 = 9.48. Interpolating from the biaxial line of Fig. 4.22, the
concrete strength coefficient is Y= 0.101. Since the column aspect ratio exceeds 2, and
rie longitudinal reinforcement ratio is 0.025, both a@ and B =1.0. The strength of the
concrete shear-resisting mechanisms, with "ce = 39 MPa, is thus, from Eq.(4.72):
Vo =a. fF, (O8A,) = 0.101V39(0.8% 2.543) = 1.283MN(289kips)
Axial Load component: The axial load is 3.82MN, and the depth of the compression
zone, 6, is 460mm. The angle of the axial load strut is thus given by tang =(0.9-0.23)/10
=0.067. From Fg.(4.73a), the shear strength resulting from the axial load component is:
V,, = 0.85P.tan ¢ =0.85%3.82x 0.067 = 0.218MN(49kips)
Transverse reinforcement component. Equation (4.76b) applies, with @= 35°. The
pitch of the hoops is s= 120mm, and hence the available strength is
Ay f(D -¢~c,)cou9) _ _314x420(1800— 460-40) cot 35
5 2 120
* = 3.214MN(723kips)
‘The total shear strength is thus:
Vy Vo 405 4Vp =1.28340.218+3.214 =4.71SMN >3.628MN (OK).
In fact, a spacing of 180mm (7.1in) would be adequate for required shear strength, but
che more critical confinemem requirement of 120mm governs, and is thus specified. Note
that the shear requirement is not added to the confinement requirement — the same
reinforcement can simultaneously provide confinement and shear strength. If the
reinforcement is stressed by shear strength requirements, it will bear against the core
concrete, providing effective confinement to the concrete. On the other hand, if the
transverse reinforcement is in tension duc to confinement, this stress can be used to
ansfer shear force across open cracks. If it is further considered that the peak shear and
confinement demands occur in orthogonal directions, it is clear that the column need
only be designed for the more critical of shear and confinement.
4.7.5. Shear Strength of Reinforced Concrete and Masonry Walls
(a) Shear Demand: The design sheat force for reinforced concrete or masonry walls is
strongly influence by capacity design influences ~ particularly higher mode effects. This is
discussed in some detail in Section 6.6, and is only briefly considered here. Figure 4.28
shows two profiles of moment up the height of a cantilever wall, with Fig.4.28(a)
corresponding to the design distribution of lateral forces, resulting from the DDBD
procedure, and Fig.4.28(b) including a possible distribution resulting in maximum feasible
base shear. In Fig.4.28(b) the base moment M” exceeds the design moment Mp, as a
result of material strengths exceeding expected values (see Sections 4.2.4 and 4.5).184 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
10
8
z | \ Design profile
£6
z
Ba
=
He
2-
o
Moment Mp Moment Mp M®
(a) Design Forces (b) Maximum Shear Response
Fig.4.28 Moment Profiles for a Cantilever Structural Wall
Critical higher mode effects result in a much more rapid decrease in moments with
height than for the design profile, and hence the overstrength base shear force, given by
the slope of the moment profile at the wall base (V’=M°/H") is significantly larger than
the design base shear force (Vp =Mp/H,). In the example shown in Fig.4.28, with M? =
Mp, the overstrength shear is V’ = 2.2Vp.
(®) Concrete Shear-Resisting Mechanisms: ‘The strength of the concrete shear-
resisting mechanisms can be determined from Section 4.7.3(a) as for columns. However,
it could be noted that at maximum shear response, the ductility demand is unlikely to be
as high as the design level, since the desiga level for displacement and ductility demand is
dominated by first mode response, while peak shear demand will be strongly influenced
by second, and perhaps third-mode response, which are unlikely to occur simultaneously
with peak first mode response. ‘The aspect ratio for the wall to be used when determining,
the factor @in Eq,(4.72a) should be based on the overstrength effective height; that is,
MVD = Hlly.
With Hanged walls, it is recommended that the effective shear area be based on the
web dimensions (in the direction considered) and that the area of any flanges be
conservatively ignored.
(©) Axial Force Component: The concepts outlined in Section 4.7.3(b) for the
contribution of axial force in shear resistance were related to a single axial force
component. In multi-storey walls, axial force, and its influence on shear strength will be
composed of a number of components; one for each floor level. Thus Eq.(4.73) shouldChapter 4. Analysis Toots for Direct Displacement-Based Design 185
be replaced by a surnmation of the effects from each floor considered separately. The
load at each floor should be considered to be applied through the geometric centroid of
the wall, and the resulting angle of the compression strut calculated separately for each
floor.
(d) Transverse Reinforcement Mechanisms: Again the recommendations of Section
4.7.3(0) can be directly used. When the wall flexural reinforcement is uniformly
distributed along the wall length, as we recommend, it will help to control the diagonal
flexural-shear cracking, and hence enhance the concrete shear-resisting mechanisms.
Also, with longer walls, the critical 35° angle of diagonal cracking is likely to intersect a
floor slab, potentially mobilizing additional shear resistance. With typical storey heights of
3m (l0fi) this can be expected for walls longer than about 2.5m (8.2f). Although this
may not be relied upon in design, it provides additional security against shear failure.
4.7.6 Response to Seismic Intensity Levels Exceeding the Design Level
When structures are designed to the damage-control limit state under a level 2
carthquake, there remains the possibility that they may be subjected to higher intensity
levels, pethaps corresponding to level 3 (see Section 2.2.2.(c)). It is clearly important that
to satisfy the life-safety performance criterion required for a Level 3 earthquake, shear
failure must not occur. The influence of increased seismic excitation on shear strength is
expressed in the proposed design approach by an increase in curvature ductility, which
results in a decrease in the steength of the concrete shear-resisting mechanisms (see
Fig4.22). Generally the inherent conservatism of the proposed shear design approach will
be sufficient to ensure safe performance under a Level 3 earthquake. However, if the
concrete shear-tesisting mechanisms provide a large portion of the total shear strength,
added conservatism in the shear design would be appropriate. This could be effected, for
example, by using the minimum levels for the parameter 2 in Fig.4.22(a).
4.8 SHEAR FLEXIBILITY OF CONCRETE MEMBERS
48.1 Computation of Shear Deformations
For slender members, with aspect ratios M/ VD > 3, shear deformation will be small,
and can be ignored. It will be recalled that in Section 4.2.7 the use of a linear elastic
curvature distribution in predicting force-displacement response was. partly to
compensate for shear deformation, It should also be noted that ignoring shear
deformation is not necessarily unconservative in a displaccment-based design
environment, since shear deformation will increase the displacement capacity
corresponding to strain-based flexural limit states.
For members of low aspect ratio, and particularly when assessing the force-
displacement characteristics of existing structures, it may be advisable to consider shear
deformation. In this context it should be pointed out chat only approximate methods for186 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
predicting shear deformations of concrete members are available, though compression-
field theory! provides a means for predicting total deformation, including shear
components, for monotonic response of members with positive stiffness. The following
simplified approach is modified from an approach developed by Miranda et all, and
assumes a bilinear flexural force-displacement approximation is utilized. Adaptation 10
the “refined” approach of Fig.4.9 is obviovs.
Calculation of shear deformation is divided into three phases of the force-
displacement response:
(a) Elastic, Prior to Shear Cracking: With the bilinear approximation to response, it is
assumed that the clastic phase is represented by constant flexural section stiffness of
Eleg = My, a7)
In the elastic range, prior to the formation of diagonal shear cracks, flexural cracking,
reduces the shear stiffness approximately in proportion to the reduction in flexural
stiffness, Thus, the shear stiffness in this phase may be approximated by:
Loe Ely
sa = ks Bp
(4.78)
ross
where the shear stiffness for an uncracked cantilever column of shear area Ax, shear
modulus Gand length His given by:
GA,
k, gross i
For solid sections, the shear area may be taken as Ay = O.87Agms, with adequate
accuracy, and the shear modulus as G = 0.43 E, for concrete and masonry structures.
Shear cracking may conservatively be assumed to occur when V=Ve, given by
Eq.(4.72), with 7= 0.29 (MPa units; 7= 3.5, psi units). Note that this corresponds to the
initial strength for the assessment model, which is used here in preference to the design
model, since a best escimate of deformation is required. ‘The displacement at onset of
diagonal cracking is thus:
Ay Volk oy (4.80)
(4.79)
(b) Elastic, After Shear Cracking: The shear stiffness for incremental displacements
after onset of diagonal shear cracking is based! on a model developed originally by
Paulayi?3"l, ‘This considers the shear Aexibitity of an equivalent strut-and-tie model,
incorporating both the compression of the diagonal strut, and the extension of the tie
representing the transverse reinforcement. The unitary shear stiffness (ie. shear stiffness
of a unit length of member) for members with transverse reinforcement perpendicular to
the axis is given byChapter 4, Analysis Tools for Direct Displacement-Based Design 187
_ p, sin’ @.cot”
sin’ @+np,
E.bd (4.81)
where Bis the angle the diagonal strut makes with the member axis, py is the area ratio of
sransverse reinforcement, m = E,/Ez is the steel/concrete modular ratio, by is the
effective width of the section, and dis the effective depth of the section.
rhere are some problems associated with the use of Eq.(4.81), particularly in applying.
it to circular sections. One general problem, however, is that for typical values of and
Pa it implies that the shear stiffness inoreases as the angle 8 reduces from 45° to 30°, This
pears counter-intuitive, as members suffering shear distress typically exhibit a
decreasing angle @as shear force increases to maximize shear strength (see Eq.(4.76),e.g)
‘That this should be accompanied by decreasing shear deformation is unlikely, and is not
supported by experimental results, so far as we are aware. Nevertheless, Eiq,(4.81) gives
reasonable prediction of shear deformation of rectangular columns when @ = 45°.
With this approximation, Eq.(4.81) can be simplified to
= OP. epg us)
0.25+np,
A further problem with Eqs(4.81 and 4,82) is the evaluation of the modular ratio. This
will increase as the flexural strength of the column is approached, resulting in a decrease
in effective shear stiffness. ‘This occurs as a result of softening of the diagonal
compression strut, Consequently, the value for m should reflect this. An average value of
n= 10 is recommended.
For circular columns, as mentioned earlier, additional problems are associated with
definition of Py by and d. The following approximate values are suggested:
AL
2°D's
=0.39p, and b, (4.83)
where Py is the volumetric ratio of transverse reinforcement (see Sec.4.2.2).
‘The shear stiffness defined by Eq.(4.82) applies for shear forces between diagonal
cracking, and nominal flexural strength. At nominal flexural strength the shear
deformation is thus:
Ay =A, + Vy Ve Wk, (4.84)
where Vy_ is the shear force corresponding to flexural strength.
(6) Ductile phase: \n the post-yield phase the concrete compression struts within the
plastic hinge region will continue to soften, and thus shear deformation will continue to
increase, Experiments on columns where shear failure does nor occur have indicated that
the shear deformation, as a fraction of total deformation remains essentially constant, or188 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
reduces slightly, as the ductility increases. Since rational computation of the shear
deformation in this phase of response is difficult, and the significance to total
deformation is typically small, it is eecommended that the shear deformation be increased
in proportion to the flexural deformation after yield, That is:
A.
A, arg (4.85)
If the “tefined” force-displacement representation of Section 4.2.7 is adopted,
Eq.(4.85) should be applied after the member has attained its nominal flexural strength.
48.2 Design Example 4.3: Shear Deformation, and Failure Displacement of a
Circular Column
‘The column of Design Examples 4.1 (Section 4.5.1) and 4.2 (Section 4.7.4) is now re~
analysed with the effective column height reduced from 10m (32.8{t) to 5m (16.46%) to
make it more shear-critical. Note that the column aspect ratio is now 5/18 = 2.8,
indicating that shear deformation probably should be considered. Longitudinal and
transverse reinforcement, and axial load remain as in the previous examples. First the
force-displacement response needs to be reassessed.
Table 4.1 Force-Displacement Data for Example 4.3 (IMN=225kips, 1m=39.37in)
Row | Moment | Curvature Shear Flexural | Shear Disp | Total Disp
QviNm) (axt)__| Force(Mtn) | Disp. (mm) | (mm) (mm)
7 00 00 00 00 0.00 0.00
2 18.87 (0100242 3.774 24.23 181 26.04
ES 19.88 0.00260 3.976 24.75, 2.95 28.70
4 2131 (0.00298 4.262 28.55, 456 33.11
5 22.36 0.00339 4472, 31.22 475 36.98
6 24.08 0.00450 4816 3.70 7.70 44.40
7 24.13 0.00571 4.826 44.24 "8.88 53.12
& 0.00695 4.946 5057 9.56 60.14
9 26.21 0.00814 5.242 56.59 10.10 66.70
10 26.72. (0.00928) 5344 (2.42 T7144 73.55
vy 27.14 (0.01038 5.428 68.00 12. 80.10
a 27.85 (0.01250 5.570 78.66 14.04 92.70
B 28.39 (0.01453 5.678 88.78 i484 104.62
14 28.89 0.01651 5.778 98.62 17.60 116.22
15 29.34, 0.01849 5.868 108.44 19.35 im
16 29.72 0.02044 5.944 118.06 21.07 139.13
i7 30.36 0.02421 6.072 136.60 24.38 160.58
8 30.84 0.03533, 6.168 142.31 24.39 167.70,Chapter 4, Analysis Tools for Direct Displacement-Based Design 189
‘Table 4.1 includes data necessary to construct the force-displacement response. The
moment and curvature are values found from moment-curvature analysis of the cokumn,
using the procedures outlined in Section 4.2, and the overstrength data corresponding to
curve B in Fig.4.20. The data in Row 2 correspond to first yield of the extreme
longitudinal reinforcement. Because the axial compression stress is low, flexural cracking
occurs at a very low moment, if the recommendation to take tension strength as zero is
adopted. Consequently, the moment-curvature response is essentially linear up to first
vield, and intermediate values have been omitted, for simplicity. The “refined” force-
displacement approach of Section 4.2.7 can thus also be assumed to be linear before first
yield. Shear force is simply found from V=M/5.
Flexural displacement: Since the yield strength for this assessment is taken as 1.34 =
546MPa (79.1ksi), the strain penetration length needs to be recalculated from the value
used in Example 4.1: Lsp = 0.022%fyedy) = 0.022%546x40 = 480.5mm (18.9in)
Plastic hinge lengur, From E.q,(4.31), Lp = 0.08x5000+480.5 = 880.5 >2x480.5
i.e. Lp =961mm (37.8in)
First yield displacement. Ftom Table 4.1, Row 2, the first yield curvature is @’y =
0.00242, ‘Thus, A’, = 0.00242x(5+0.481)2/3 m = 24.23mm (0.95in).
Subsequent flexural displacement. From Eq.(4.37), the flexural displacement for
moments higher than first yield are given by
A= wy theloo ha
The displacements resulting from Eq,(4.86) are listed in Table 4.1 in Column 5
(Flexural Disp. (mm)), and the force vs. flexural displacement is plotted in Fig.4.29 by the
dashed line labelled Flex. Disp.
)
= +{¢- 0.00242 - 0.961x5000mm (4.56)
18.87)
Shear displacements prior to shear cracking: Following, the recommendations of
Sec. 4.8.1, the shear stiffness in the elastic range of flexural response is given by Eq.(4.79).
This requires the ratio of effective to gross flexural stiffness to be calculated. The
effective stiffness can be calculated from the moment-curvature data at first yield as
= M,/9', =18.87/0,00242 = 7798MNm?
Now:
Togs = AD! 164 = mi 8)" 164 =
E, = 5000V39 =31.2GPa-
.515m*, and from Eq.(4.43)190 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Assessment strength
Shear Force (MN)
0 50 100 150 206 250
Displacement (mm)
Fig.4.29 Force-Displacement Response, and Shear Strength, for Example 4.3
(IMN = 224.8kips; Imm = 0.0394in)
Hence:
El.eg | El yy, = 7798 ((3 1,200 0.515) = 0.485
‘The shear force at diagonal cracking is, from Eq.(4.72) with a= B= 1.0:
0.29/39 x 0.8% 2.543 = 3.68MN(827kips)
The clastic shear displacement at diagonal cracking is thus, with G=0.43x31.2
=13.42GPa and Ay = 0.87%2.543=2.21 2m
VH El goose 3.68 5000
GA, Elg — 13,420*2.2120.485
= 1.28mm(0.05in)
Note that in this case diagonal shear cracking and flexural yield occur almost at the same
shear force.
Shear displacement after diagonal cracking: Using the recommendations of
Eq.(4.81) for circular columns, the effective area ratio of transverse reinforcement is
taken as py = 0.39p, = 0.39X0.006=0.00234, and b, = d= 0.8x1.8 =1.44m. Substituting
into Eq.(4.82):Chapter 4, Analysis Tools for Direct Displacement-Based Design 191
= 0.250, gp g = 0.25%0.00234
= bd =. 200,000 1.44? = 885 MN
0.25 +np, 0.25-+10%0.00234
This is the unitary stiffness (stiffness for a unit column height). ‘The column shear
stiffness is thus
Koster = Bge | H = 885/5 = 177 MN Im
‘The incremental shear displacement for shear forces higher then the diagonal cracking
shear force Ve is thus:
Displacements resulting from this equation are added to the elastic shear displacement
at diagonal cracking (see Eq.(4.84)), and listed in Column 6 of Table 4.1. These apply only
up to the nominal flexural strength. In accordance with Eq.(4.85) the shear displacements
have been increased pro-rata with the flexural displacements from Row 9 (which
corresponds to nominal flexural capacity based on the enhanced yield strength) onwards.
Ic will be seen that shear displacements are about 18% of the flexural displacements.
Column 7 of Table 4.1 lists the total (flexural-+shear) displacements, which are plotted by
the solid line, marked “Tota! Disp.” in Fig4.29.
Shear strength. Again, it is assessed that the column could be subjected to ductile
response in orthogonal directions, and thus the biaxial data of Fig. 4.22(b) apply.
Initial shear strength: At a displacement corresponding to fly = 1, the assessed shear
strength will be as follows
Ve = 0.29V390.8%2.543
82(0,9-0.23)/5
Vs = 3.887MN (from Ex. 4.2, but with 630°)
Total
o8MN
1512MN
3.887MN
.083MN (1817kips)
Shear strength at full ductility: \t curvature ductilities of ly > 13, the assessed
strength of the concrete shear-resisting mechanisms decreases to:
Ve = 0,05V39x0.8%2.543 = 0.635MN (143kips),
and the total shear strength is thus 4.034MN (1832kips). The displacements at which
these to limits apply have been calculated from the biaxial approximation to the force-
displacement response, conservatively ignoring shear displacement.
‘The assessment shear strength envelope is plotted in Fig.4.29 by the solid line. It will
be seen that it intersects with the total displacement curve at 162mm (6.4in), which is very
close to the assessed flexural displacement limit. Thus the assessment is that shear failure192 Priestley, Calvi and Kowalsky. Displacement-Based Sei
jic Design of Structures
is unlikely to occur before the displacement capacity is reached (since the force~
displacement response is based on maximum feasible flexural reinforcement strength).
Note that for a new design, however, a more conservative approach is required, using the
design strength components of Sec.4.7.3 and a shear strength reduction factor of 0.85,
‘The design strength envelope based on this approach is also shown in Fig.4.28 (by the
dash-dot line), and indicates a safe design displacement of only 76mm (3.0in), about 50%
of the displacement that would be permitted in an assessment situation. Note that
satisfying the conservatism of the design approach would be easy to provide (the hoop
spacing in the potential plastic hinge would be reduced to 75mm (2.95 in)) but the
consequence of excessive conservatism in assessment of an existing column could mean
expensive retrofit measures.
4.9 ANALYSIS TOOLS FOR DESIGN RESPONSE VERIFICATION
4.9.1 Introduction
Direct displacement-based design (DDBD) is a simple method for determining the
required strength of plastic hinges to satisfy a specified performance limit state, defined
by strain or drift limits. Combined with capacity-design requirements defined in the
structure-related sections of this book, it provides a complete seismic design approach for
comparatively simple and regular structures. It may also be used to provide preliminary
estimates for member strengths for structures which do not fir the “simple and regular”
criteria, In these latter cases, and in cases where the structure has special importance, due
to function, or cost, design verification by additional analysis will be required.
It has already been established (sce Section 1,3), that clastic modal analysis should be
considered unsuitable as a design tool, for @ number of reasons. Similarly, elastic modal
analysis will generally be considered unsuitable for design verification, though the
deficiencies cited related to determination of initial stiffness no longer apply, since it will
be possible to use reasonable estimates of stiffness based on actual provided strength (sce
Section 4.4). However, the inability of elastic modal analysis methods to model the
variation of flexural stiffness with axial force remains. ‘This leaves two candidates for
acceptable analysis methods: time history analysis, and static inelastic (pushover) analysis.
Aspects relating to these methods are discussed in the following sections.
4.9.2 Inelastic Time-History Analysis for Response Verification
(@) Elastic or Inelastic Time History Analysis? Although some design procedures
accept elastic time-history analysis (ETHA) as an analysis approach, our view is thar there
is little value in its use for design verification. Ic suffers from most of the deficiencies
apparent in clastic modal analysis, and cannot represent the differences resviting from
structural systems with different hysteretic characteristics. With one-dimensional (line)
element modelling, computational times are generally not large, even with inclastic timeChapter 4, Analysis Tools for Direct Displacement-Based Design 193,
history analysis (THA), and current lap-top computers are capable of analysing complex
structures under three-dimensional response. As a consequence of these considerations,
the following discussion is limited to ITHA.
Inelastic time-history analysis provides the most accurate method for verifying that
inelastic deformations and rotations satisfy the design limits, and also for determining the
higher mode effects, which are needed for defining the required design strengths of
capacity-protected members.
Despite the above comments relating to the suitability of lap top computers for
analysis, and by inference, the viability of ITHA in the design office environment, i¢ must
bbe recognized thar a large number of subjective modelling decisions will generally be
needed, and itis essential that the importance of these choices be properly understood by
the analyst, who should have appropriate experience in ITHA, and knowledge of
material behaviour before using it for design verification. ‘The following attempts to
provide some guidance in these choices.
(&) Degree of Sophistication in Element Modelling. The degrce of sophistication in
modelling, and the computational time, will largely be dictated by the choice of elements
used in member modelling: line, fibre or solid elements. The choice should be dictated by
Einstein’s maxim, that analysis should be as simple as possible, but no simpler. In general
this will result in a decision to use line elements.
(i) Line elements: Beams and columns in stractural frames, bridge piers, wharf piles and
other members whose connection to adjacent members can be idealized as a point
connection are normally represented by line members. With suitable moment-rotation
hysteresis characteristics, the non-linear flexural response characteristics can be modelled
with considerable accuracy, and simplified representations of axial and shear
deformations can also be incheded. Non-linear axial force /moment interaction envelopes
enable the strength characteristics of columns to be represented with adequate accuracy,
Normally, bilinear moment rotation properties will be used to represent the strength
envelope, with a lumped-plasticity representation. For conerete and masonry members,
elastic and post-yield stiffness characteristics should be based on moment-curvature
analysis of the member (sce Section 4.2.1), with the elastic stiffness represented by the
secant stiffness to the “first yield” point of the response (Fig.4.6). Thus, using the
formulation of Section 4.2.6, the effective moment of inertia will be:
1,-fe
"Eb, Eg,
(4.87)
‘The post-yield stiffness will also be found from the moment-curvature response of the
member considered, and a hysteretic rule will be chosen characteristic of the member
type and material used. Hysteretic rules are discussed in more detail in Section 4.9.2(0194 Priestley, Calvi and Kowalsky, Displacement-Based Seismic Design of Structures
In Section 4.4 it was noted that the clastic stiffness of reinforced concrete and (by
analogy) reinforced masonry columns and walls is a function of the axial force. Thus in a
frame, where the column axial force varies as a result of seismic response (¢.g. columns A
and C in Fig4.30(a)), the elastic stiffness should also vary. Although many inelastic
analysis computer codes can model the axial force/moment strength interaction, few
have the capability of modelling the axial force/stiffness interaction. An exception is the
2D/3D code “Ruaumoko"e", A “student” version of Ruaumoko, with restricted
capabilities is included on the CD provided with this book. Lack of ability to model the
axial force/stiffness relationship can result in significant errors in columa member forces,
particularly in the elastic range of response, but is unlikely to result in significant errors in
displacements and inelastic rotations, nor in significant errors in steel frame member
forces.
With frame analysis, careful consideration of modelling the beam-column joint
stiffness is needed. A common error is to model the joint as effectively rigid. However,
due to strain penetration, and joint shear deformation, the joint region is far from rigid,
and may contribute as much as 30% to total lateral deformation|"!, As illustrated in Fig.
4.30(b), 2 minimum acceptable representation of the joint region will be to use lumped-
plasticity line members to. represent the beam beoween the column faces, with a linear
clastic portion of clastic stiffness equal to the beam clastic stiffness, from the joint
centroid to the column face. If the column is to be modelled by an inelastic element, a
Location of
lumped
plasticity
plasticity line
element
elastic
(a) Column Seismic Axial Forces (b) Joint Representation
Fig. 4.30 Modelling a Regular Frame with Line ElementsChapter 4. Analysis Tools for Direct Displacement-Based Design 195
similar treatment will be provided. However, if the design philosophy is to keep the
column elastic, it may be adequate to use a simple column element from joint centroid to
joint centroid, and eliminate the nodes at the top and bottom of the beam, Elastic
members will not be appropriate for ground-floor columns, where column-base plastic
‘hinges are expected to form, Some time-history analysis codes (e.g. Ruaumokol") enable
special joint deformation elements to be located between the nodes at the joint
boundaries. This will normally only be appropriate when it is desired to model joint
failure,
The effects of strain penetration from critical sections into supporting members,
briefly discussed above in relation to beam/column joints, must also be considered at the
base of frame columns, and at top and bottom of bridge piers by additional elastic
members. Numerical simulation of this is discussed in detail in Section 4.2.7. It is also
important that the effects of foundation flexibility, including the foundation structure and
she supporting soil be adequately modelled. There is little point in carrying out a
sophisticated ITHA if a flexible foundation is unrealistically modelled as rigid. Vertical,
wanslational and rotational stiffness may be modelled by elastic of inelastic springs.
Specific advice is provided in chapters dealing with specific structural types.
(i) Fibre elements: \ number of ITHA codes (e.g, SeismoStruct!*t!) for concrete structures
are based on representing the cross-section of linear members by a number of fibres,
separately representing the concrete and reinforcing steel, as illustrated in Fig.4.31. The
length of the member is divided into a number of segments, with each segment
represented by fibre elements, and with the sections delimiting the segments following
che Navier-Bernoulli approximation that plane sections remain plane. Each of the fibres
represents an area of concrete, or a longitudinal reinforcing bat, and is given appropriate
material properties to model reversed loading (ic. arbitrary tension/compression
histories). Typically, separate material rules are used for the unconfined cover concrete,
and the confined core concrete, and in advanced formulations (e.g. SeismoStruct), the
stress-strain properties of the confined concrete are automatically generated. Material
stress is normally assumed constant between integration points along the fibre segment.
The advantages of this mose sophisticated representation of linear members are:
* No prio: moment-curvature analysis of members is needed.
* The hysteretic response is defined by the material properties, and hence does not
need to be defined.
* The influence of varying axial force on strength and stiffiness is directly modelled.
Simulation of biaxial loading, is as straightforward as uniaxial loading, with
interaction between flexural strength in orthogonal directions being directly
computed.
* Member post-peak strength reduction resulting from material strain-softening or
failure can be directly modelled.
‘There are, however, also draw-backs to modelling using fibre-elements. With current
formulations:196
Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Reinforcement
fibres ele a Cover fibres
Core fibres
Fig.4.31 Section Modelling with Fibres
Fibre elements model only flexural response. Shear strength and shear
deformation are generally not modelled,
The interaction between flexural ductility and shear strength (see Section 4.7.3) is
not modelled
Shear deformation in joint regions must be modelled by special non-fibre
elements.
Integration of the segment curvatures to obtain rotations and displacements
suffers from the same problems identified in Section 4.2.6 for direct integration
of curvature from linear-member analysis. Ta some extent this can be obviated
by adopting a critical segmene fength equal to the plastic hinge length computed
in accordance with Eq.(4.30).
If the moment in a segment is determined at the segment centroid, then the
capacity of the structure may be over-estimated, since the moment at the critical
section will exceed that at the segment centroid, which is the location where
strength will be determined. Again this can be corrected in the analysis, by
adjusting the location of critical segment boundaries (e.g. the beam-end at a
beam-column interface) such that the integration gauss points coincide with the
true boundary
Strain-penetration into foundations etc, requires special treatment.
Properties for the confined core concrete may have to be manually defined
However, in some fibre analysis programs, as with some moment-curvature
programs, the stress-strain characteristics of the core concrete are directly
determined from the transverse reinforcement.
Because of the large number of fibre elements needed to fully model a complex
structure, computer time for ITHA can be very large with current computing
power ~ as much as 24 hours for a single run. This makes design verificationChapter 4, Analysis Tools for Direct Displacement-Based Design 197
impractical if (say) seven records are to be run, ‘The problem may be particularly
severe if 3-D response is needed.
Ir should be emphasised that in fibre analysis it is not essential, nor even desirable that
all members (or even all segments of a member) be modelled by fibres. Joint rotations
and foundation flexibility effects can be modelled by elastic or inelastic springs, with the
appropriate degrees of freedom, and members expected to remain clastic with essentially
constant stiffness may be modelled by clastic line elements. In fact, given the major
current drawback of excessive computational time, this simplification should be adopted
wherever possible. The fibre program “SeismoStruct” is provided on the attached CD.
1it) Three-dimensional solid elements: Taree-dimensional solid elements are used rather
infrequently in ITHA, are of value only when modelling of complex intersections
between members is required, and should only be attempted by highly specialized
organizations, Computational effort is considerable, and in our view, existing
formulations do not adequately represent bond-slip and anchorage of reinforcement,
confinement effects, and the behaviour of unconfined or confined concrete under
reversed flexure and shear loading where separate patterns of intersecting cracks may
develop. The cost of ITHA using inelastic 3-D elements is typically high, and the
accuracy of the results is difficult to assess. For these reasons, we believe design
verification should be confined to line and fibre elements at the current state of the att.
(c) Two-dimensional or Three-dimensional Structural Representation?: Ail
structures are three-dimensional, However, in many cases it will be reasonable to
independently consider the seismic response in the principal directions, and hence to
provide simpler two-dimensional representations of the structure, Examples are straight
bridges, and symmetrical buildings where response in the orthogonal directions are likely
ro be essentially independent. Two-dimensional structural representation is generally
significantly easier to develop and to interpret the results from than is the case with 3-D
representation, Further, when columns are expected to respond inelastically in orthogonal
directions (bridges, building column bases), 3-D analyses will require special hysteretic
rules modelling the interaction of strength and ductility in orthogonal directions, unless
Abre elements are used, Such rules are not readily available, and the current state-of-the-
art is to model the ductility effects independently, though a biaxial strength envelope may
‘be available, depending on the computer program used.
When significant torsional response is possible, as in wharves (see Chapter 12), and
some buildings (Chapters 5, 6 and 7), then 3-D representation of the structural response
‘is necessary. However, as, for example, discussed in Chapter 12 for wharves, it is
sometimes possible t capture the salient features of the 31D response with 2-D
structural modelling.
(d) Strength Interaction Modelling: \s discussed in Section 4.9.2(b) above, many
ITHA programs provide means for modelling the axial-force/flexural-strength
interaction, either ditectly, as in the case of fibre elements, or as specified input, as in the198
stley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
case of line elements. However, as was discussed in Section 4.7.3 there is also an
interaction between flexural ductility and shear strength with concrete structures. ‘This is
much more difficult to represent in ITHA, and we are not aware of any general-purpose
ITHA program currently capable of modelling this interaction. For design verification
this is not generally a problem, as the design philosophy will be to avoid shear failure, and
the results from the ITHA in terms of member ductility and shear demand can be used
in conjunction with the shear strength model of Section 4.7.3 to design the transverse
reinforcement.
Where ITHA js used to assess the safety of an existing structure, the inability «0
model the shear-strength/flexural-ductility interaction envelope, and the subsequent rapid
reduction in shear strength once the envelope has been reached is of more serious
concern, Although it may be possible to identify the level of seismic intensity
corresponding to just reaching the envelope, the increment of intensity required to cause
collapse is currently beyond the capabilities of existing commercially available programs.
(¢) Mass Discreeization: Masses will normally be lamped at nodes, implying a degree
of mass discretization, In some computer programs the discretization will be carried out
automatically from values of element mass per unit length provided as input; in orher
programs the nodal masses will be directly input.
(i) Frame buildings: Referring again to Fig, 4.30(a), nodal masses in 2-D representations
will normally be provided at the intersections of all beams and columns. ‘This will enable
realistic values for seismic axial forces in the beams to be determined. In some cases these
axial forces can have a significant effect on beam moment capacities. However, in most
cases it will be admissible ro lump the entire floor mass at one representative node. In
the case represented in Fig.4,30(a), this would clearly be the node defined by the central
column. Note that tributary column mass is generally lumped at the floor levels above
and below the column considered. Note also that increasing the number of nodal masses
will increase the number of significant elastic modes, some of which will have very low
periods. This can result in computational problems, and so excessive refinement of the
mass distribution should be avoided.
When 3-D modelling and 2-D or 3-D seismic input is adopted, it is important that the
mass torsional inertia is correctly represented. Referring to the plan simulation of a
rectangular floor slab shown in Fig.4.32, a minimum of four nodes is required to fully
represent the mass torsional inertia, Assuming that the mass is uniformly distributed, the
distances of the mass node points are at the torsional radii of gyration respectively
(L/N12) and (B/V12) from the floor centroid in the longitudinal and transverse
directions.
(ii) Wall buildings: Similar considerations apply for wall buildings. Generally floor masses
will be concentrated at the centreline of the wall or walls. Again, excessive refinement in
the assumptions of the mass distribution should be avoidedChapter 4. Analysis Tools for Direct Displacement-Based Design 199
i L/NI2—ke— B12
Fig.4.32 Plan Distribution of Mass for To:
Floor Slab
nal Inertia Modelling of a Building
(a) Multi-span Bridge
ky
m
ky
Deflection Pier Geometry Mass and Stiffness
(b) Bearing-supported Superstructure
Fig.4.33 Mass Discretization for Bridge Inelastic Time-History AnalysisChapter 4. Analysis Tools for Direct Displacement-Based Design 201
shown with the elastic and initial inelastic cycles shown by dotted lines, and subsequent
cycles shown by solid lines. Where the initial and subsequent cycles coincide they are
shown by solid lines. ‘The elasto-plastic rule of Fig.4.34(a) has been widely used in ITHA
in the past, but is of comparatively restricted applicability. It can be considered an
approximation to the flexural response of steel beams and columns, provided the critical
sections can be considered compact (no flange buckling) under inelastic response.
However, the reduction of stiffness on unloading, (the Bauschinger effect) which causes
rounding of the hysteretic response and 2 reduction of energy absorption of about 20%
compared with elasto-plastic response, is not modelled. Elasto-plastic response is
appropriate for seismic isolation systems using planar coulomb friction isolation sliders
‘eg. PTFE/stainless stcel flat bearings). For these elements the initial stiffness normally
results from the structural elements, with the inelastic portion defined by essentially
rigid/ perfectly plastic behaviour of the slider
The bilinear hysteretic response of Fig, 4.33(b) has a post-yield stiffness of 20% of the
initial stiffness (ic. r = 0.2), though the actual value of the post-yield stiffness will vary
depending on the specific application, Compact steel sections are better represented by
bilinear response with 0.02<1<0.05 than by elasto-plastic response, which is a special case
of bilinear response with r= 0. The high value of r= 0.2 in Fig 4.34(b) is representative
of seismic isolation provided by elastomeric bearings or friction-pendulum systems, with
the value of r depending on the initial structural stiffness, and the isolation system
characteristics.
Figures 4.34(¢) and 4,34(d) are based on the modified Takeda hysteresis rulel™, and
represent a range of hysteresis shapes appropriate for reinforced concrete and reinforced
masonry structures. The modified Takeda rules are characterised by unloading and
loading stiffnesses thar are significantly lower than the initial “elastic” stiffnesses. The
“thin” hysteretic response of Pig. 4.34(0) is appropriate for inelastic members with
significant axial load, such as building columns, bridge piers, walls and piles, and is a
special case of the more general “fat” Takeda rule shown in Fig.4.34(d). For the thin
Takeda role the unloading stiffness is defined as:
k, =k,
(4.88)
where Lis the displicement ductility at the initiation of unloading, On the initial cycle the
rule reloads to the elastic yield point in the reversed direction, but on subsequent cycles it
reloads to the previous maximum force-displacement point, This is a special case of the
“fae” rule, with B= 0.
The “fat” Takeda rule of Fig.4.34(d) incorporates the “thin” rule as a special case, as
noted above, and has an unloading stiffness defined by
k, =k (4.89)202 Priestley, Cal
ind Kowalsky. Displacement-Based Seismic Design of Scructures
(©) “Thin” Modified Takeda (r.c.column) (4) “Fat” Modified Takeda (rc. beam)
() Flag (Hybrid prestressed concrete)
( Gapping (soit)
Fig. 4.34 A Selection of Hysteresis Rules Appropriate for Inelastic Time-History
Analysis
where 0<@<0,5, and [1 is again the displacement ductility. A value of @=0 implies
unloading with the initial stiffness. The rule reloads 10 a point defined by BAp (see Fig
4.34(@)) from the previous peak displacement in the direction of loading considered,
where Ap is the plastic displacement. For well-detailed reinforc
of @=0.3, and B=0.6 are generally considered to be appropriate.
Note that with the Takeda rules, the initial stiffness should be the value applying at
first yield, taking into account the reduction in effective stiffness resulting from flexural
and shear cracking (Eq.(4.86))
Figure 4.34(e) describes a flag-shaped hysteresis rule suitable for modelling the
response of hybrid prestressed members. In hybrid members (see Section 5.11.2) the
prestressing steel is unbonded, resulting in a non-linear clastic force-displacement
d concrete beams, valuesChapter 4, Analysis Tools for Direct Displacement-Based Design 208
characteristic. The energy dissipation in hybrid systems, shown in Fig.4.53(e) results from
vield of special mild-steel bars which contribute to both flexural strength and energy
dissipation, Further details are given in Chapter 5.
Finally, Fig. 4.34(6) shows a hysteresis rule appropriate for a spring representing non-
linear soil response, such as might occur in the soil near the top of a laterally loaded pile.
After an initial cycle involving inelastic response of the soil spring, the soil separates from
the pile, leaving a gap. On reloading in the same direction there will be no soil reaction
unl this gap is closed. Typically inelastic soil response has a comparatively high post-
sield stiffness, as illustrated in Fig.4.34(6.
(g) Elastic Damping Modelling: It is common to specify a level of elastic damping,
in ITHA to represent damping in the initial stages of response, before hysteretic damping
is activated. This is normally specified as a percentage, typically 5%, of critical damping,
‘There are a number of ways this damping could be defined, as discussed subsequently,
but the principal difference is whether the damping force is related to the initial or
cangent stiffness, In the hysteresis rules of Fig.4.34, the initial stiffness is ky in all cases,
bur the tangent stiffness, rh; (see Fig 4.34) reduces to a lower value once the
displacements exceed yield.
‘Typically research papers reporting results on single-degree-of-freedom (SDOF)
ITHA state that 5% clastic damping was used, without clarifying whether this has been
related to the initial or tangent stiffness, With mult-degree-of-freedom (MDOF)
analyses, the situation is often further confused by the adoption of Rayleigh damping,
which is a combination of mass-proportional and stiffness-propostional damping. It is
our understanding that many analysts consider the choice of the initial clastic damping
model to be rather insignificant for either SDOF or MDOF inelastic analyses, as the
effects are expected to be masked by the much greater energy dissipation associated with
hysteretic response. This is despite evidence by others (e.g. {O4]) that the choice of initial
damping model between a constant damping matrix and tangent-stiffness proportional
damping maurix could be significant, particularly for short=period structures,
The difference between initial-stiffness and tangent stiffness damping is discussed
with reference to Fig. 4.34(d). For SDOF systems, the constant value of the damping
coefficient is determined with respect to the initial vibration frequency, @= {k,/m , the
initial loading stiffness kj and a specified fraction of critical damping, ¢
c= 2magg =2E Jk, (4.90)
‘The damping force at any instant is thus
Fy =x (4.91)
where 4 is the instantaneous relative velocity.204 Priestley, Calvi and Kowalsky.Displacement-Based Seismic Design of Structures
With tangent-stiffness damping the damping coefficient is proportional to the
instantaneous value of the stiffness and it is updated whenever the stiffness changes.
‘Thus the damping coefficient given by Eq.(4.90) is multiplied by Vkd/Ai, where hy is the
instantaneous tangent stiffness. With reference to Fig4.34(d), the damping coefficient
will equal that associated with the initial-stiffness value only in the initial elastic response.
After first yield, the damping coefficient will be referenced to the post-yield, the
unloading or the reloading stiffness, With the case of elasto-plastic response, the damping
force will be zero while the structure deforms along a yield plateau.
‘There are three main reasons for incorporating elastic damping in ITHA:
The assumption of linear elastic response at force-levels less than yield: Many
hysteretic rules, including all those shown in Fig.4.34 make this assumption, and.
therefore do not represent the nonlinearity, and hence hysteretic damping within
the elastic range for concrete and masonry structures, unless additional elastic
damping is provided.
¢ Foundation damping: Soil flexibility, nonlinearity and radiation damping are not
normally incorporated in structural time-history analyses, and may provide
additional damping to the structural response.
© Non-structural damping: Hysteretic response of non-structural elements, and
relative movement between structural and non-structural elements in a building
may result in an effective additional damping force.
Discussing these reasons in turn, it is noed that hysteretic rules are generally
calibrated to experimental strucrural data in the inelastic phase of response. Therefore
additional elastic damping should not be used in the post-yield state to represent
structural response except when the structure is unloading and reloading elastically. If the
hysteretic rule models the elastic range nonlinearly (as is the case for fibre-element
modelling) then no additional damping should be used in ITHA for structural
representation, It is thus clear that the elastic damping of hysteretic models which have a
linear representation of the elastic range, and which hence do not dissipate energy by
hysteretic action at low force levels would be best modelled with tangent-stiffness
proportional damping, since the elastic damping force will greatly reduce when the
stiffness drops to the post-yield level. It should, however, be noted that when the post-
yield stiffness is significant, the elastic damping will still be overestimated. This is
particularly important for hysteretic rules such as the modified Takeda degrading stiffness
rule which has comparatively high stiffness in post-yield cycles.
If the structure deforms with perfect plasticity, then foundation forces will remain
constant in the structural post-yield stage, and foundation damping will cease, It is thus
clear that the effects of foundation damping in SDOF analysis are best represented by
tangent stiffness related to the structural response, unless the foundation response is
separately modelled by springs and dashpots.
It is conceivable, though unlikely, that the non-structural damping force is velocity-
dependent rather than strength-limited, and hence a constant damping coefficient may
bea reasonable approximation for the portion of “elastic” damping that is attributable toChapter 4. Analysis Tools for Direct Displacement-Based Design 205
non-structural forces. There are two possible contributions to non-structural damping
that should be considered separately:
‘Energy dissipation due to hysteretic response of the non-structural elements
* Energy dissipation due to sliding (relative movement) between non-structural and
structural elements.
For a modern frame building, separation between structural and non-structural
elements is required, and hence they should not contribute significantly to damping,
Further, even if not separated, the lateral strength of all non-structural elements is likely
to be less than 5% of the structural lateral strength (unless the non-structural elements are
masonry infill), If we assume 10% viscous damping in these elements, an upper bound of
about 0.5% equivalent viscous damping related to the structural response seems
reasonable. Non-structural elements are unlikely to play a significant role in the response
of bridges.
Sliding will normally involve a frictional force, with the value dependent on the
friction coefficient, and the weight of the non-structural element. Unless the non-
structural elements are masonry, the friction force is likely to be negligible. It should be
noted that it is probable that so-called non-structural masonry infill initially contsibutes,
more significantly to strength, stiffness, and damping than is the case with (eg)
lightweight panitions. However, it is known thar the strength degrades rapidly for drift
levels > 0.5% (which is generally less than structural yield drift), The damping force is
also likely to degrade rapidly.
Ieis thus recommended that for modern buildings with separated or lightweight non-
structural elements, clastic damping should be modelled by tangent stiffness proportional
damping. The effects of non-structural masonry infill should be modelled by separate
structural elements with severely degrading strength and stiffness - not by increased
viscous damping,
Te is instructive in determining the influence of alternative clastic damping models t0
consider the steady-state, harmonic response of an inelastic SDOF oscillator subjected to
constant sinusoidal excitation. This enables direct comparison between hysteretic and
dastic damping energy, and also between clastic damping energy using @ constant
damping coefficient and tangent-stifiness proportional damping models. To this ead,
Fig4.35 shows response of a simple SDOF oscillaror with inital period of 0.5 sec and a
“thin” modified Takeda hysteresis rule (Fig.4.34(0)) subjected 10 10 seconds of a 1.0 Hz
forcing function. The steady-state response of the pier corresponded to # displacement
ductility of about 7.7 — at the upper limit of reasonable ductile response.
Results for the stabilised loops, ignoring the transitory first three seconds of response
are plotted in Fig. 4.35(a) (initial-stiffness proportional damping) and Fig, 4.35(b)
(tangent-stiffness proportional damping). In each case the hysteretic response associated
with nonlinear structural response is plotted on the lef, and the elastic damping force-
displacement response is plotted to the right. The areas inside the loops indicate the
relative energy absorption. For the case with initial-stiffness proportional damping, the206 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
energy absorbed by elastic damping is approximately 83% of the structural hysteretic
energy dissipation, despite the high ductility level. This might be surprising when it is
considered that the elastic damping corresponds to 5% of critical damping, while the
hysteretic damping is equivalent to about 20% of critical damping, ‘This apparent
anomaly is due to the different reference stiffness used. The elastic damping is related to
oo a 5
sw 4 =m
g g
3 g
Boo Eo
:
® 1000 4 & .1000 4
aio + am |, :
a1 ° a a ° oa
Displacetent én) Displacetnent x)
(a) Analysis with Initial Stiffness Damping
ot 5 200 =
| =o
g zg
ee 5° =
000 9 1000
ane : 20 1, :
a ° a an ° Mi
Displacement m) Displacement ()
(b) Analysis with Tangent Stiffness Damping
Fig.4.35 Steady-state Inelastic response of an SDOF oscillator with “Thin”
Modified Takeda hysteresisChapter 4. Analysis Toots for Direct Displacement-Based Design 207
the initial stiffness, whereas the hysteretic damping is related to the secant stifiness to
maximum response.
When the elastic damping is tangent-stiffness proportional, as we believe to be most
appropriate for structural response, the clastic damping energy is greatly reduced, as can
be seen by comparing the upper and lower right-hand plots of Fig. 435. In the lower
plot, the reduction in damping force corresponding to the stiffness change is clearly
visible, In this case, the area of the elastic damping loop is only about 15% of the
sxeuctural hysteretic energy dissipation
Analyses of SDOF systems subjected to real earthquake records show that the
significance of the elastic damping model is not just limited to steady-state response
Figure 4.36 shows a typical comparison of the displacement response for SDOF
oscillators with initial stiffness and tangent stiffness elastic damping, in this example the
El Centro 1940 NS record (amplimde scaled by 1.5) has been used, the initial period was
0.5 seconds, a Takeda hysteretic rule with second slope stiffness of 5% was adopted, and
the force-reduction factor was approximately 4. The peak displacement for the tangent
stiffness elastic damping case is 44% larger than for the initial-stiffness damping case,
indicating a very significant influence.
A selection of results from a series of analyses of oscillators with different initial
periods between 0.25 sec and 2.0 sec, subjected to NTC32MI spectrum-compatible
0s
oa 4
0.05 5
Displacement (m)
i
0.05 5
-o 4
0 4 8 2 6 2»
‘Time (seconds)
Fig 4.36 Response of an SDOF Oscillator to 1.5xEl Centro 1940,NS (T:208 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
accelerograms is shown in Figure 4.37, which plots the ratio of peak inelastic
displacement response to peak clastic displacement response. Elastic analyses, using 5%
damping (note that the choice of damping model js irrelevant for efastic analysis) were
first carried out, then inelastic analyses where either initial-stiffness or tangent-stiffness
5% damping were specified, with yield strengths based on force-reduction factors of
R=2, 4 and 6 based on the average elastic response peak force. Modified Takeda
(representing concrete response), Bilinear (approximating steel) and Flag (representing
hybrid prestressed precast concrete) hysteresis rules were considered.
The second-slope stiffness ratio for the “Thin” Takeda (Fig.4.34(c)) and bilinear rules
for these analyses was r= 0.002 (the minimum value considered), but similar results were
obtained for r= 0.05 (the maximum considered). The flag hysteresis results in Fig.4.36
were based on the minimum additional damping considered in the hybrid precast
modelling (8 =0.35, sce Fig.4.34(e)). Note that in some cases the displacement ratios a T
25 sec. have not been plotted as they exceed the range included by the graph axes.
From examination of Figure 4.37 it will be noted that there is a significant difference
between the response of the initial-stiffness and tangent-stiffness (identified as IS and TS
respectively) models, that this difference is rather independent of initial period, for
T>0.5 seconds, that the difference increases with force-reduction factor, and is
dependent on the hysteretic rule assumed. It will also be aoted that though the “equal
displacement” approximation (represented by a displacement ratio of 1.0 in Figure 4.37)
is reasonable for initial stiffness damping and initial periods greater chan "= 1,0 seconds,
it is significantly non-conservative for tangent-stiffness clastic damping.
It is thus recommended that for modern buildings with separated or lightweight non-
structural elements, elastic damping should be modelled by tangent-stiffness proportional
damping, Note, however, that the specification of tangent-stiffness Rayleigh proportional
damping for MDOF, multi-storey buildings will not have the desired effect, since most
of the clastic damping in the critical first mode will be mass-proportional, which is
constant with inelastic action. Consider the basic form for the fraction of critical damping
in the Rayleigh damping model:
(4.92)
where @ and B are the coefficients associated with mass proportional and stiffness
proportional damping respectively, and @, is the circular frequency. If we specify the
same value (say 5%) for & at two different frequencies, where the higher one is times
the lower (fundamental) frequency, then:
(4.90)
and°
2
3
g
=
Displacement Ratio (AAtastic)
Displacement Ratio (A/Actastic)
Displacement Ratio (A/Actassc)
os
a
16
12
08
2a
16
os
“Tools for Direct Displacement-Based Design 209
0 of 08 12 16 2 0 94 08 12 16 2 0 o4 08 12 16 2
Period (seconds) Period (seconds) ‘Period (seconds)
(a) Takeda (concrete) hysteresis
STEELE nn nan RE LR een i a een
0 of 08 12 16 2 G 04 o8 12 16 2 o 04 08 12 16 2
Period (seconds) ‘Period (seconds) Period (seconds)
(b) Bilinear hysteresis
T m4 TT 9 TIT
0 04 08 12 36 2 0 of 08 12 16 2 0 O84 08 12 16 2
Period (seconds) Period (seconds) Period (seconds)
()Fiag Hysteresis
Figure 4.37 Response of SDOF Oscillators to ATC32 Spectrum-Compatible
Accelerograms210 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
f )
iat (4.91)
eo!
2
Subtracting lig. (4.91) from xtimes Eq (4.90):
}. 4.92)
and hence the damping @/20 associated with mass proportional damping in the first
mode is x/(k-+1), while the damping attributed to stiffness proportional damping in
the first mode is €(x+1)
Consider the case where we specify 5% damping at 7; = 1.5 sec and T= 0.3 sec.
Hence = 5. Then, even if we specify tangent-stiffness proportional Rayleigh damping,
only 0.83% is stiffness proportional in the critical first mode, while 4.17% is mass
proportional, and hence acts én an identical manner to initial-stiffness damping when the
structure responds inelastically
It is clear from the above discussion that it will be difficule to obtain the correct
simulation of tangent-stiffness clastic damping, when Rayleigh damping is specified. It
will always be preferable to specify pure tangent-stiffness damping, at least for the
fundamental mode. This option may not be available with commercially available ITHA
codes, and an acceptable alternative is to specify an artificially low damping coefficient,
&*in the fundamental mode. Based on the discussion on significance of elastic damping,
t0 DDBD, presented in Section 3.4.3, this value should be approximately:
eal »
0.10
uid +ru-r)
A more complete discussion of the issues related to elastic damping is available in
(P16, and G2|
(4.93)
(h) Choice, Number, and Character of Accelerograms for Design Verification:
() Number of records: Two alternatives are generally defined by codes (c.g; X3,X4) for the
number of accelerograms to be used in design verification. The first involves using three
spectrum-compatible records, with the design response being caken as the maximum
from the three records, for the given response parameter (displacement, shear force etc)
investigated. The second uses a minimum of seven spectrum compatible records, with
the average value being adopted for the response parameter considered. Because of the
simplicity of ITHA with modern computing power, the latter approach is now almost
always adopted, and there appears to be a tendency to increase the number of records
above the minimum of seven, to ensure a more representative average.Chapter 4, Analysis Tools for Disect Displacement-Based Design 2M
(ii) Seketion of records: There are three basic choices for the means of obtaining spectrum-
compatible accelerograms:
* Amplitude scaling of acceleration records from real earthquakes to provide a
“best fit” to the design spectrum over the period range of interest.
© Generating artificial spectrum-compatible records using special purpose
programs.
‘* Manipulating existing “real” records to march the design spectrum over the full
range of periods.
When the records are obtained by amplitude scaling of existing records, the scatter
between records is likely to be large, and hence a larger number of records might be
needed to obiain a reliable average. Care has to be exercised in selecting the period range
over which the spectrum matching is obtained. In Fig. 4.38, which shows the matching
referred to the 5% damping elastic displacement spectrum, the matching has been
reasonably achieved over the period range encompassing the first thrce elastic periods,
which might be considered adequate. However, it is seen that at periods longer than the
fundamental period, displacements of the scaled spectrum are significantly lower than the
desiga spectrum, It is important to match the spectrum for a period range that includes
the period shift expected as the structure responds inelastically. This is shown in Fig.4.38
for a displacement ductility of ft = 3. In this example, the scaled displacement spectrum
becomes increasingly deficient in displacement demand, compared with the design
spectrum, as the period increases above the clastic period, resulting in an unconservative
estimate of the inelastic displacement from the time-history analysis. OF course, this
argument applies to a single record. Uf a large number of records are used such that the
average of the displacement spectra over the full range from elastic to inelastic period
matches the design spectrum, valid average results can be expected, Unfortunately this is
unlikely to be the case for longer period structures, since the large majority of records
used in the analyses tend to have peak spectral displacements at periods from 1.0-2.0
seconds. In this case there is likely to be a consistent deficit in displacement demand at
the degraded (inelastic) fundamental period, regardless of how many records are used.
Thus conclusions about structural response for structures with expected displacement
ductilities of (say) [= 4 and elastic periods of T >0.75 seconds can be expected to be
suspect, unless the earthquake records are very carefully chosen. Note that though the
argument above has referred to code spectrum matching, it also applies to more general
studies related to investigation of the relationship between clastic and inelastic
displacement.
‘The second alternative involves artificially generated accelerograrns, using programs
such as SIMQKEI, ‘These can be matched to the design spectrum for the full period
range with comparatively small error, and can be given the general character of an initial
segment of increasing intensity, a duration of essentially constant intensity, and a final
segment of reducing intensity. A lesser qumber of records are required to obtain a
meaningful average using artificial records. An objection commonly voiced about artificial
records is that they ate too severe, in that real records are not spectrum-compatible over212 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Design :
Speci 37” | ror
Displacement
Specteal
matching
Ts T, Period
Fig.4.38 Displacement Spectrum Matching for Inelastic Time-History Analysis
the full period range, and that artificial records typically have a longer duration than real
records. The first of these arguments cannot be accepted, since the larger number of real
secords needed to obtain a full spectrum matching should produce essential identical
results, and the artificial records can be considered as a more efficient means of obtaining
the same results. The second objection — excessive duration ~ is unlikely to be of concern
when the analyses are being carried out for verification of a new design. When the
analysis is carried out on an existing structure which is expected to degrade in strength
under the design spectrum, then duration may be significant. However, it should be
recognized that few ITHA programs based on line elements have the ability to accurately
model strength degradation, though this is sometimes available with fibre elements.
Recently, the third option, where existing recorded accelerograms are manipulated to
obtain full spectrum matching, has become more common. This method has the
advantage over pure artificial records that the essential character of the original record is
preserved. Thus, records that conform to the type of source characteristics expected (e.g,
strike/slip, subduction, near-field forward directivity etc.) can be selected. This is
important, since the spectrum matching will normally be done at 5% damping, and the
characteristics at different levels of effective damping will depend on the source
characteristics and distance from the fault plane. However, to obtain the required
spectrum matching, the duration of the record typically has to be extended, opening the
method to the same objection as directed against artificial records. Figure 4.39 shows
typical spectral matching of acceleration and displacement spectra based on manipulating
seven real accelerograms.
(iii) Multi-component Accelerograms: When structural response is likely to be influenced by 2-
D or -D effects — that is, when the response in orthogonal directions cannot reasonablyChapter 4. Analysis Tools for Direct Displacement-Based Design
(a) Acceleration Spectra
40
(c) Displacement Spectra
Fig. 4.39 5% Spectral Matching Using Manipulated Real Recordsl#!!
213214 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
be de-coupled, the accelerograms will have to include two or three orthogonal
components. The generation of the characteristics of the various components needs
careful consideration. Examination of recorded three-component accelerograms reveals
that the two horizontal components have different spectral shapes and intensities, but it
does not seem that there is a consistent difference when decomposed to fault-normal and
fault-parallel directions. Thus choice of a predominant, or principal direction of attack
does not seem feasible at this stage of knowledge. Consequently site-specific design
spectra are generally generated using the average of faul-normal and fault-parallel
actenuation relationships. In the event that a code-specified spectrum is adopted, this will
almost certainly be independent of orientation.
‘The vertical component typically has lower peak spectral acceleration intensity — a
value of 0.67 times the peak horizontal spectral acceleration intensity is often quoted,
though this appears to depend on distance from the fault, hypocentral depth, and ground
material — and a systematic difference in spectral shape, with peak acceleration intensity
typically occusring at a lower period than for the horizontal components, and wid
intensity reducing with period more rapidly than with the horizontal components!"
Recent near-field accclerograms of shallow earthquakes have indicated vertical
accelerations that have in many cases exceeded the peak horizontal components.
The above discussion raises the question of how the orthogonal components of real
records should be scaled, In the following it is assumed that no preferential direction for
the design spectrum has been identified. There are a number of possibilities:
‘© Systematically rotate the axes of the owo horizontal records, to determine the
principal direction, and generate new major and minor principal-dizection
accelerograms. Scale the major principal record to the design spectrum, and use
the same scaling factor for the minor horizontal and the vertical components.
* Scale the larger of the records to the design spectrum, without determining the
principal direction, and use this scaling factor for the other two components.
* Scale both horizontal records independently to the design spectrum, and scale
the vertical spectrum to the vertical design spectrum.
It might appear that the first, or possibly the second alternative is the most
satisfactory. However, without a knowledge of the direction in which che major
component should be applied with respect to the structure axes (which is the current
state of the science), multiple analyses would be required, with the direction of the
horizontal components rotated by (say) 15° berween successive analyses to capture the
maximum response. This approach also has the disadvantage of not recognizing how the
design spectrum is obtained. As noted above, attenuation relationships used to generate
site-specific design spectra normally are based on the average of faul-normal and fault-
parallel response. Thus no specific principal direction is implied by the design spectrum.
It thus appears that both records should be scaled to the design spectrum. Since the
records will have low cross-correlation, it is expected that the elastic response
displacement in any direction will be similar, when the two horizontal components are
applied simultaneously in any two orthogonal directions. This is illustrated in Fig. 4.40,
whete the response displacements for different bearings are compared with the designChapter 4. Analysis Tools for Direct Displacement-Based Design 218
Dir-2
Dirt “1
1-Second Period, 1.5-Second Period, 2-Second Period,
Ave.=7” vs. target=7.05” — Ave.=11.7" vs target=12.1” — Ave,=15.2” vs target=14.8”
3-Second Period 4-Second Period, 5-Second Period,
Ave.=23.6" vs target=24.2” Ave.=27.2" vs target=26.6" _Ave.=25.5” vs target=27.8"
Fig.4.40 Response Displacements in Different Bearings for Two-component
Spectrum-Compatible Records Applied in the NS and EW directions, Compared
with the Spectrum Targets!"
values for different specified elastic periods, and with the two spectrum-compatible
components applied in the NS and EW directions. The data, and records, relate to a study
for the Port of Los Angeles(@!! and ate the average of seven spectrum-compatible
accelerogram pairs. Tt will be seen that good agreement with the circular design response
is obtained for all except T=Ssec, which was beyond the period range of interest.
This approach appears more compatible with the way in which the design spectra are
generated, and has the great advantage that a single analysis related to any chosen set of
axes is sufficient, rather than the multiple analyses required by the first ovo alternatives. It
should, however, be applied with some caution, When excitation in one direction
produces significant displacement response in the orthogonal direction, ic would appear
that use of evo spectrum-compatible records will produce higher estimates of peak
displacement than use of scaled major and minor principal-direction accelerograms
rotated to a series of possible orientations, as discussed above.
@ Averaging Results from Multiple Analyses: \t might be considered that averaging
the results from a set of (say) seven spectrum-compatible cecords or record pairs is216 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Maxima Displacements
Record | A Positive | A Negative
I 6.341 ~6.347
2 4.891 27.945
3 8.948 “3.179
Concrete 4 4.486 ~6.543
Reinforcement} Compression 5 6.148 4.569
strain 6 6.198 4.710,
T| 4.667 77.032.
Average | _ 5.954 -5.761
Fig.4.41 Bridge Pier Response Displacements from Seven Spectrum-Compatible
‘Accelerograms
straightforward. In fact some careful consideration of rather simple probabilistic aspects,
is required. Consider the response displacements of the simple bridge pier shown in Fig,
4.41, subjected to uniaxial seismic excitation. The table indicates the positive and negative
displacement maxima obtained from each of seven spectrum-compatible accelerograms,
and the averages for the positive and negative directions. The maximum of the positive
and negative displacement for cach record is highlighted with bold text. What is the
correct average response displacement? If the absolute value of the response
displacement is of interest, then the average of the highlighted values should be used,
giving a value of 7.023, which is 18% and 22% higher than the averages in the positive
and negative directions respectively. However the absolute maximum. displacement,
regardless of sign, is unlikely to be a significant design parameter. If longitudinal response
is considered, the displacement may be required to determine whether a movement joint
closes up, causing impact, or opens up sufficiently to cause unseating, In this case, since
the critical displacements will be different in the different directions, the sign of the
displacement is important
Under transverse (or longitudinal) response, the displacements may be critical to
determine whether stecl and concrete strain limits are exceeded in the plastic hinge
region, The critical locations of the pier for positive displacement are indicated in Fig.
4.41, For these two locations and design parameters (concrete compression strain, and
reinforcement tension strain) it would clearly be inappropriate to include the negative
displacements in the averaging, since these provide strains of the incorrect signs in the
cctitical region.
‘The argument supporting the inclusion of the larger negative displacements is to note
that the polarity of the excitation for a given record is arbitrary, and hence the
displacement signs ate arbitrary, indicating that averaging the larger of the positive andChapter 4, Analysis Tools for Direct Displacement-Based Design 217
negative displacements is correct. This argument is, however, invalid, as it implies
considering the response for both negative and positive polarities, (which essentially
implies consideration of 14 rather than 7 accelerograms), sclecting the response of the
seven largest, and discarding the response of the seven smallest. From a probabilistic
viewpoint it would be equally valid to select the seven smallest and discard the seven
largest. In fact it is obvious that no selection can be justified, and if both positive and
negative polarities are considered, the results of all 14 records should be averaged. Using,
the data of Fig. 4.41, this would result in a design displacement of 5.858, 17% lower than
the value from averaging the peak displacement magnitudes
Averaging the positive and negative displacements would only be valid if the structure
has symmetric strength and stiffness characteristics in the opposite directions. For non-
symmetric structures, the positive and negative displacements should be separately
averaged. Alternatively, additional analyses could be run, reversing the polarity of the
records to provide 14 valid values for averaging.
The illogicality of selecting the maximum of the positive and negative displacements
becomes even mote apparent when 2D or 3-D excitation is considered. If the critical
polarity of each component is to be adopted, then four possible combinations of the 2D
components (+ve/+ve; -ve/—ve, +ve/-ve; -ve/+ve) and eight possible combinations of
the 3-D components would need to be considered.
‘There are, however, additional problems with determining the correct average for
response under multi-axial excitation. Consider the case of the simple wharf segment
shown in plan view in Fig, 4.42 which has considerable eccentricity between the centre of
mass and the centre of strength (effective stiffness), as a result of short piers on the
landward edge and long piers on the seaward edge, as illustrated in Fig, 12.1(a). Because
of the significant torsional response under longitudinal (X direction) excitation the
teansverse and longitudinal response cannot be decoupled, and simultaneous excitation in
the orthogonal (X and Y) directions must be modelled.
1 ay
(Critical Pile 1 __Landward edge
OF} Centre of Strength
Oy cin icici
{> Centre of Mass
A Seaward edge
Fig 4.42, Peak Corner Displacements for Critical Corner Pile of a Wharf
(Plan View)218 stley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
‘The critical design parameter will be the average displacement from the seven (or
more) pairs of accelerograms (record pairs) of a corner pile. As with the previous example
of the bridge pier, the average displacement is of interest because it can be directly
correlated to the pile limit strains
Output from the ITHA will be in the form of X and Y displacements (as well as
member forces etc). Normally the peak values Amer,x and Amacy from the analysis will also
be listed in a summary table. However, it is not possible to combine these vectorially to
obrain che peak response displacement for the record under consideration, since Amuse
and Anasy are unlikely 10 occur at the same time. Instead, at each time interval 4, the
vectorial displacement A, must be calculated according to:
——
A= yA, +415) (4.93)
and the peak displacement for the record determined from the full range of A; thus
obtained. This peak displacement will occur in a specific direction. In Fig, 4.42, che
critical displacements and directions for the first four of the record pairs are indicated by
vectors. It will be clear that these cannot be averaged as scalars to obtain the design
response displacement, as they occur in different directions. The correct method of
averaging becomes apparent when one considers the condition that defines the limit
strains. Let ws first consider the peak response in a specified direction (say 15° fiom the X
axis). The average displacement response in this direction will define whether the limit
strains on the diagonal defined by the direction are satisfactory. ‘To obtain this average,
the displacement in this direction must be obtained for cach record pait, and for each
time step. The full time-history of response displacememss in this direction is searched for
each record pair to obtain the maximum value, and the average of the maxima in this
direction from the seven (or more) record pairs is found.
This procedure must be carried out for a series of directions around the full 360
Normally directions at 15° intervals provide sufficient accuracy. The average response
displacements in the different directions are then searched to find the ctitical direction,
and thus the critical displacement.
It is noted that this procedure represents a departure from what has been customary
practice in the past, but we believe it represents the only logical interpretation of che
averaging process for response under a number of accelerograms.
4.9.3. Non-Linear Static (Pushover) Analysis
Pushover analysis involves an inelastic analysis of the structure considered, under a
gradually increasing vector of forces or displacements, representing the expected pattern
Of inertia forces or response displacements in the structure. It has the ability to track the
formation, and plastic rotation of plastic hinges in the structure, and hence can be of
value in design verification. There are, however, limitations to the utility of the method.Chapter 4. Analysis Tools for Direct Displacement-Based Design 219
First, it can not be directly used to determine the overall displacement demand on the
seructure under the design seismic intensity. This must be obtained from some other
means, such as modal analysis, since what is obtained is essentially a generalised force-
displacemem cesponse of the structure, Given knowledge of the displacement demand,
the pushover analysis can be carried out until the displacement at some characteristic
location reaches the assessed demand, and hence the local inelastic strains and structucal
cirifts can be determined. Thus assumptions made in the direct displacement-based design
approach about the deflected shape can be tested, but not the absolute value of the
displacements,
A further limitation of the pushover method is that at the current stage of
development pushover analysis in commercially available analysis codes is generally
restricted to modelling response of a single mode, generally the fundamental mode.
Dynamic amplification of drifts, moments and shear forces due to higher mode effects
cannot be accurately modelled with single-mode pushover analysis. Considerable research
effort has been expended in recent years towards developing multi-mode pushover
analysis, In our view these have not yet reached maturity, and require excessive
computational effort. This severely restricts the applicability of the method to design
verification,
A final limitation is that with a uni-directional push, the hysteretic characteristics of
the structure cannot be included for straight design verification, Thus, a structure with
nonlinear elastic hysteretic characteristics will have the same pushover response of a
bilinear elasto-plastic system with the same backbone response. In some computer
pushover approaches!*!, however, cyclic pushover analyses caa be carried out. These can
be used to determine the energy dissipation characteristics of sub-assemblages of the
structure, enabling them to be modelled with a reduced number of degrees of freedom in
ITHA, or in the design process. ‘This is of considerable value in soil-structure interaction
problems. An application to the response of wharves is presented in Chapter 12.
In carrying out a pushover analysis, a decision must be made as to whether to contro
the response by an imposed force-vector or by an imposed displacement-vector, Both
approaches have limitations. The force approach tends to become unstable as maximum
resistance is approached, and cannot follow response when strength degrades. It also
results in a force vector that is generally incompatible with the generated displacements,
since in a single-mode pushover, the force vector should be proportional to the product
of mass and displacement at each level, In some approaches!\"), more than one shape of
force vector is specified (typically corresponding +o a linear or constant displacement with
height), and che results from the worst case is taken. ‘This implies considerable doubt as
to the accuracy of the results.
The displacement approae!
response in the post-peak strength region, However, spurious local forces can develop
when applied to MDOF structures, and the specification of a displacement vector can
inhibit soft-storey failure mechanisms. tn both force-controlled and displacement-
controlled cases, care must be taken in applying an appropriate force or displacement
vector that represents the expected nse. Recently developed pushover
is very suitable for SDOF systems, since it can follow
modal respe20 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
algorithms, termed “adaptive pushovers” modify the applied force or displacement vector
incrementally to model the changing deflecuon shape resulting from the structural
analysis", and should be used where possible
It is pointed out that data preparation and interpretation for a pushover analysis is
very similar to that required for ITHA, and with current computing capacity, the
increased time 10 execute an ITHA compared with a pushover analysis is of lite
consequence, unless fibre elements are used. Considering the limitations of pushover
analysis, outlined above, we see limited applications for the approach, at this stage of
development of the method, in design verification studies, beyond the specific examples
noted above. It should be noted however, that the increased complexity of fibre elements
which creates problems in excessive time of analysis for ITHA, can enhance the accuracy
of pushover analyses without involving excessive analysis time5
FRAME BUILDINGS
5.1 INTRODUCTION
This chapter builds on the basic material provided in Chapter 3 relating to the
fundamentals of dixect displacement-based seismic design (DDBD), to provide a rather
complete procedure for the seismic design of buildings whose primary lateral force-
zesisting systems is comprised of frames. ‘The relevant equations from Chapter 3 are first
summarized, and extended, where appropriate, Information on elastically responding
frame buildings is provided, since in many regions of low to moderate seismicity frame
buildings can be expected to respond elastically to the design-level seismic intensity.
Simplified methods of analysis under the design seismic force-vector are presented,
since it is shown that traditional elastic analyses are invalid for frames that respond
inelastically. Details on the necessary strength margins of columns to satisfy capacity
design requirements are considered in some detail. Design examples are used throughout
«o illustrate the complete design process, for both steel and concrete frame buildings
5.2. REVIEW OF BASIC DDBD PROCESS FOR FRAME BUILDINGS
5.2.1 SDOF Representation of MDOF Frame
Chapter 3 introduced the fundamentals of DDBD with respect to Fig.3.1, which is
reproduced here for convenience as Fig.5.1. The first stage of the design process is the
representation of the multi-degree-of-freedom (MDOF) structure by an equivalent
single-degree-of-freedom (SDOF) structure modelling the first inelastic mode of
response (Fig.5.1(a)). The following equations were developed
(a) Design Storey D.
related to a normalised inelastic mode shape , where /= 1 to mare the storeys, and to
he displacement A, of the critical storey by the relationship
acements: ‘The design floor displacements of the frame are
\
a,=6/4e
J
221222 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
(a) SDOF Simulation (b) Effective Stiffness Ke
025 = 05, =
Lo
Damping Ratio, &
1
0.05 +}
Opes
Displacement Ductility Period (seconds)
(c) Equivalent damping vs. ductility (A) Design Displacement Spectra
Fig.5.1_ Fundamentals of Direct Displacement-Based Design
where the normalized inelastic mode shape depends on the height, Hi, and roof height
Ay, according to the following relationships:
forn <4:
(5.2b)
for n >4:224 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
is generally sufficiently accurate to assume a linear yield displacement profile for the
purpose of estimating ductility demand, and hence the yield displacement is given by:
»=O,-H, (68)
@ Equivalent Viscous Damping: From Section 3.4.3, the equivalent viscous
damping of the substitute structure for frames can be conservatively related to the design
displacement ductility demand (see Fig.5.1(©) as follow
Reinforced Concrete Frames: &, = nas+0.ses( 4-1) (6.92)
un
u-l
Steel Frames: oq = 9.05 + 0.577 -| —— (5.9b)
un
Alternatively, the distributed ductility approach of Eq,(3.37) can be used.
(@@) Effective Period of Substitute Structure: ‘The eflective period at peak
displacement response is found from the displacement spectra set (Fig.5.1(@)), entering
with the design displacement and determining the period, Ts, corresponding to the
calculated equivalent viscous damping.
(h) Effective Stitiness of Substitute Structure: With reference tw Fig.5.1(b) and
Section 3.2, the effective stiffness at maximum displacement response of the substitute
structure, F/Ay is
K, =4z°m,/
6.10)
@ Design Base Shear Force: Again with reference to Fig.5.1(b) and Section 3.2, the
design base shear force for the MDOF structure is found from the substitute structure:
F =Vagg = K Ay G.11)
5.2.2 Design Actions for MDOF Structure from SDOF Base Shear Force
(@) Distribution of Base Shear Force to Floor Levels: ‘he base shear force from
q,6.11) is distributed to the floor levels in proportion to the product of mass and
displacement as:
F= VruuclmA,)(omA,) (6.12)
iatChapter 5. Frame Buildings 225
(6) Design Moments for Plastic Hinges: The building is then analysed under the
force vector represented by Fq.(5.12) to determine the required flexural strength at the
plastic hinge locations. As discussed in Section 3.7 it is unnecessarily conservative to add
the full gravity-load moments to these seismic moments. So doing, as well as increasing
the cost of the structure would reduce the seismic response displacements to below the
intended design level. The recommendation in Section 3.7 is to detail the beam plastic
hinges for the lower of (i) the seismic moments, and (ii) the factored gravity-lo:
noments. This is discussed further in Section 5.6.4
Analysis of the building under the lateral force vector requires the adoption of
member stiffness appropriate for the expected member ductility level, as discussed in
Section 3.5.7. However, an alternative and simpler approach to determining the member
forces is presented in Section 5.5.2.
5.2.3 Design Inelastic Displacement Mechanism for Frames.
The desired mechanism of inelastic deformation for frames involves the formation of
flexural plastic hinges at the ends of the beams, except, possibly, at roof level, combined
with column-base plastic hinges, and columa-top hinges if the roof-level beams do not
hinge. This mechanism, showa in Fig.5.2(a) where beam hinges, rather than column
hinges form at the roof level provides the greatest possible number of locations to
ismic energy, and also results in plastic hinge rorations that are very nearly
ey drift.
dissipate
equal to the inelastic stor
Ap ‘Ap
(a) Beam-Sway (b) Column-Sway
Fig.5.2. Mechanisms for Frame Inelastic Response226 Priestley, Calvi and Kowalsky, Displacement-Based Seismic Design of Stcuctutes
Capacity design measures, outlined in Sections 3.10 and 4.7 are required to ensure that
this, and only this, inelastic mechanism can develop. Column flexural strengths at
locations other than the base or top must be set sufficiently high to ensure that column
hinges rather than beam hinges do not form, since this could result in a column sway
mechanism with high plastic rotation demand on the column hinges!!l, as illustrated in
Fig.5.2(b). The two alternative mechanisms in Fig,5-2 have the same plastic drift Ap at the
roof fevel, but it will be immediately seen that the plastic rotation of the hinges in the
column sway mechanism is approximately m times that of the beam-sway mechanism,
where 17 is the number of storeys. It will also be appreciated that a single column failure
could result in total building collapse, whereas failure of a single beam is unlikely to be
critical. Column-sway (also known as soft-storey) mechanisms have been one of the most
common forms of failure of frame buildings in earthquakes!”*L
Similarly, since shear failure is brittle, resulting in strength loss and potential
catastrophic failure, the shear strength of both beams and columns must be set
sufficiently high to ensure that shear failure cannot occur. Capacity design measures for
frames are discussed further in Section 5.8.
5.3. YIELD DISPLACEMENTS OF FRAMES.
5.3.1 Influence on Design Ductility Demand
Frames are inherently flexible structures. Consider Eq.(5.7a) which gives the yield
drift of a reinforced concrete frame, If we take typical values for beam span and depth of
Ly = 6 m (19.7 ft), and hy =600 mm (23.6 in), and a reinforcement yield strength of fre
400 MPa, (58 ksi), implying a yield strain of 6,
drift of
0,002, then Eq.(5.7a) results in a yield
6.0
6, =0.5x0.002x— = 0.01 (6.13)
0.6
Note that in many countries higher yield stresses are common, and hence yield drifts
would be proportionately larger.
Most seismic design codes set drift limits corresponding to a damage-control limit
state (see Section 3.3.2) in the range 0,02 to 0.025, to limit non-structural damage, Design
ductility or behaviour factors for frames are often set as high as 5 to 8 (see Table 1.2). It
is apparent, however, that if the yield drift is of the order of 0.01, then the maximum
placement ductility demand at the non-structural drift limit will be in the
2 to 25, and hence the structural ductility limits will almost never govern,
structural dis
order of
This will be apparent in examples later in this chapter,
5.3.2 Elastically Responding Frames
‘A second consequence of the high yield drift of frames is that tall frames can beChapter 5. Frame Buildings 227
expected to respond elastically if the design seismic intensity is low or moderate, since the
vield displacement is likely to exceed the maximum displacement demand corresponding,
ro elastic (eg, 5%) damping at the start of the constant-displacement plateau of the
displacement spectrurn (see Fig.5.1(d), e.g.). Aspects which will influence this are
© Yield drift (ie. frame dimensions and reinforcement yield strength)
* Magnitude of earthquake
* Distance from faule rupture
¢ Number of storeys.
‘The tentative information given in Section 2.2.2.(b) relating to prediction of
displacement spectra can be used to gain some insight into the probability that a given
building will respond elastically co an earthquake of specified magnitude occurring on a
known fault, Since it is more common to express design seismicity in terms of peak
ground acceleration (PGA), we reformulate the information of Chapter 2 in terms of
PGA as follows:
We assume a typical elastic spectral acceleration shape for firm ground, with a
constant velocity slope starting at Ty = 0.5 seconds, Note that the assumption of a
constant velocity slope, which is implicit in most seismic codes, is compatible with the
linear displacement spectra of Fig.5.1, We also make the common assumption that the
plateau acceleration is 2.5 times the effective peak ground acceleration PGA. With this
information, and the relationship between corner period and magnitude given by Eq.(2.3)
it is possible to directly relate the clastic acceleration spectrum and the displacement
spectrum (see Fig.5.3), and calculate the maximum elastic response displacement. ‘The
ordinates of the acceleration spectrum for 0.5< TXTe sec are:
Gy) = 2.5PGA-T, (T =1.25-PGAIT 6.14)
In accordance with Fq.(2.3) the corner period Te is taken as:
T. =14+2.5(M, —5.7) — seconds (5.15)
Making the usual assumption of sinusoidal relationship between peak refative
displacement and peak pseudo-acceleration, the corner clastic displacement is related to
the PGA by:
Tl. T2125 PGA
an ag
6.16)
c
Substituting for Te from Eiq.(5.15) and simplifying:
Aye = 0.031X PGA(+2.5(M,, ~5.7)) 6.17)228 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Acceleration a(r) (g)
Displacement Ar) (m)
i T=0.
T=T,
T T T T T
Period 7 (seconds) Period T (seconds)
(a) 5% Acceleration Spectrum (b) 5% Displacement Spectrum
Fig.5.3, Design Elastic Acceleration and Displacement Response Spectra (no
scale)
Note that the data used co generate the elastic displacement spectra in Section 2.2.2
imply that PGA is inversely proportional to distance from the fault plane for distances
greater than 10 km. As a consequence, high PGA’s are possible for moderate seismic
magnitudes My, provided the distance is small enough.
Equation (3.17) can be plotted as a function of PGA and My to provide the
maximum displacement that can be expected at the effective height of the subst
structure (Eq.(5.5)). This can be compared with the yield displacement given by Eqs
and (5.8) to determine for a given structure, PGA and My, whether response is likely 10
be elastic or ductile. A typical comparison is displayed in Fig.5.4 for a reinforced concrete
frame building, based on the following structural assumptions:
© flexural reinforcement yield strength fe = 400 MPa (58ksi)
* Storey height constant at 3.5
m (11.3 f)
© Beam aspect ratio Li/hty = 10 (see 2q.5.7))
Effective height = 0.7%, (Fh, = coral building height =3.52 m (11.5 fi)
Yield displacement profile is linear.
‘These assumptions are the same as used to derive Fiq.(5.13) and hence the yield drift is
0.01. ‘The yield displacement at the effective height is thus:
A
, = 8,.(0.7x3.5n) =0.01%0.7X3.5n=0.0245n (mm) 6.18)
where mis the number of storeys.Chapter 5. Frame Buildings 29
165
24 storeys
Displacement (m)
20 storeys
16 stories
22 storeys
8 storeys
4 storeys
6 64 68 7.2 16 8
Moment Magnitude, My
Fig.5.4 Comparison of Peak Elastic Spectral Displacement for 5% Damping as a
function of PGA and Moment Magnitude, with R.C. Frame Yield Displacements
In Fig,54, the relationship between peak spectral displacement, peak ground
acceleration and moment magnitude given by Eq.(5.17) is plotted as a series of sloping
ines identified by PGA. The structural yield displacements for frames between 4 and 24
eys based on Eq.(5.18) are plotted as dashed lines. If the yield displacement exceeds
the he peak clastic spectral displacement, then the response is expected to be elastic.
‘The results indicate that if the design PGA is 0.2g, and the causative earthquake is Mu
< 6.6, then all frames of 8 storeys or higher, with similar 10, or more flexible than, the
characteristics of those on which Eq.(5.18) is based, will respond elastically. Twenty-four
story frames in a zone with PGA = 0.3g will respond elastically fer all My <8, and in a
zone with PGA = 0.5g will still respond elastically unless My > 6.9.
These results should be taken as indicative only, since they depend on the structural
assumptions, and also the assumptions used co derive the displacement spectra. For the
taller structures it is also possible that inelastic response could be developed in higher
modes than the fundamental mode of vibration, thus requiring the consideration of
capacity-design effects. With these provisos, however, it will be seen that elastic response
of taller structures is probable, particularly in regions, such as Europe, where design
carthquake magnitudes tend to be moderate (ic. My < 6.5). For such structures,
simplified seismic design criteria would be appropriate,230 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
5.3.3. Yield Displacement of Irregular Frames
Tt has long been recognized that good seismic performance is more likely 10 be
achieved when the structure is “regular”. Regular, in the context of frame buildings
normally means that bay widths are equal, there are no vertical offsets, and the structure
has no designed torsional eccentricity. Often the criteria of regularity will not be met, and
in DDBD two issues will reed consideration: the estimation of the yield displacement (so
that the design ductility, and heace the equivalent viscous damping can be determined),
and the distribution of the design base shear to different elements of an irregular
. In this section we consider the former of these two considerations.
structur
Fv, ul] I Yan Bia Displacement
EVa2 E Vag
(@) Irregular Frame (b) Overturning Moment
Fig.5.5 Seismic Response of an Irregular Frame
is irregular in that it has a short central bay with longer outer
ed in Section 4.4.6, it was
The frame in Fig.
bays. Ieis recalled that when the yield drift of frames was discu
noted that Eq,(5.7) was found to apply for exterior, interior, and nvo-way beam-columa
subassemablages. It thus follows that the yield drift of the different bays of Fig, 5.5 can be
calculatéd, with sufficient accuracy, independently of each other. From Eq,(5.7) it is clear
that the beams in the outer bays will have yield drifts that are greater than the yield drift
of the central bay. This is illustrated in Fig.5.5(b), where the bay contribution 10
overturning moment is plotted against displacement at the effective height of the
substitute structure. From F.q,(5.7(a)) the yield drifts are:
6, =0.5e, 2 8
. 6.19)
AyChapter 5. Frame Buildings 231
‘Thus if the beams in the different bays have the same depths, the yield drifts will be
proportional to the span lengths. As is shown in Fig.5.5(b), the system yield displacement
3s found by combining the moment /displacemeat response of the individual bays to form
an equivalent bilinear response. In this example, if Myand Mz are the contributions to the
overturning moment from an outer and inner bay respectively, chen the total overturning
moment and system yield displacement are:
Moras =2M, + M, (5.202)
2M0,.+M,0,» Un (00)
2M,+M,
Equation (5.20b) requires that the ratio of moment contribution to the overturning
moment of the various bays, My/Mz needs co be known before the yield displacement,
and hence the ductility and equivalent viscous damping can be determined. Note that the
absolute values of Myand Mare not needed. As will be seen shortly, a rational decision
will be to design both the short and long beams at a given level for the same moment
capacity. For generality we assume different positive and negative moment capacities of
Mune and Mae respectively. The seismic moments at full mechanism development are
indicated in Fig.5.5(a) at the third floor. Beam seismic shears in the short and long spans
will thus be in inverse proportion to the span lengths:
Mye+M.
Vg) = Y,
Ly, ® Ly
621)
For the development of a full seismic mechanism, the seismic axial forces induced in
each of the columns by the beams of the oter and inner bays are & Vg) and ¥ Var
respectively. Ignoring the column-base moments as a relatively small proportion of the
oral overturning capacity, the contributions of the outer and inner frames to the system
overturning capacity are thus:
Menu: LM # Mow) (6.21a)
a a
DMoeit Macs) (5.2tb)
a
That is, the bays contribute equally co the overturning capacity, regardless of the beam:
length. This simplifies calculation of the effective yield displacement.232 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Note that the decision to make the beam moment capacities of the inner and outer
bays equal, though rational, is not essential; any ratio may be assumed, and the effective
yield displacement calculated. Ic will be noted, however, that elastically calculated seismic
moments will be higher in the short spans than the longer spans, while gravity load
moments will be largest in the long spans, reinforcing the rationality of the abov
approach. It should also be noted that if elastically calculated seismic moments are much
larger than gravity moments, as would be the case for tube frames in regions of high
seismicity, then conventional design would require the moment capacity of the short
beams to be increased above that for the long span (refer to the relative stiffness in
Fig.5.3(b)). ‘This is another example of the irrationality of design based on initial stiffness
since increasing the strength of the inner beams, either by increasing reinforcement
content or beam depth (or both) would further increase the relative stiffness of these
beams, initially attracting higher seismic moments.
Two examples of frame buildings with vertical irregularity provided by setbacks are
shown in Fig.5.6, The symmetrical layout of Pig.5.6(a) tequites no special consideration in
design, unless the storey heights vary significantly, since the ratio of storey mass to
stiffness will remain essentially the same up the height of the building as for a building
without setbacks. The unsymmetrical layout of Fig.5.6(b) also has no influence on the
design process in the direction displayed, unless there is also a setback in the direction
perpendicular to the frame displayed. In the pexpendicular direction, the eccentricity of
the upper portion of the fame will result in a torsional response in the lower storeys of
the building, which will need consideration using the techniques developed in Section 3.7,
(a) Symmetrical layout (b) Unsymmetrical Layout
Fig.5.6 Frame Buildings with Vertical Irregularity
Frames with irregular storey heights also require little special consideration in the
DDBD approach to determine required base shear strength, and distribution of lateral
forces to the different storey levels. However, the assumed displacement profile definedChapter 5, Frame Buildings 233
by Eq.(5.2) will be increasingly inaccurate as the variation in storey heights increases, and
should be checked by analyses after a preliminary design is completed, using the
techniques outlined in Section 4.9.
5.3.4 Design Example 5.1: Yield Displacement and Damping of an Irregular
Frame
regular Building,
ays: 4.5m, 7.5m, 4.5m (14.8, 24.6, 14.8 fi)
First storey: 4.5m (14.8f0), others 3.5m
(us
Concrete: Pay = 30MPa (4.35ksi)
Reinforcement: f = 4S0MPa (65.3ksi)
> [email protected] | Masses: Level 1: = 65 tonnes (143 kips)
= 60 tonnes (132 kips)
70 tonnes (154 kips)
G
+ 4.5m |e 7.5m->| 4.5m
Fig.5.7. Reinforced Concrete Frame fot Design Example 5.1
The twelve-storey reinforced concrete frame illustrated in Fig.5.7 is irregular in the bay
lengths. It also has vertical irregularity, in that the first storey is significantly higher tha
the upper storeys, as is often the case. The building consists of four identical frames, and
ses contributing to inertial response of each frame are indicated in Fig.5.7. Roof
mass exceeds that at lower floors due to equipment and water storage located at roof
level. It is required to determine the design ductility level, and equivalent viscous
damping, for a DDBD of the frame, based on a design drift limit of 0.
Solution: To determine the system ductility, and hence equivalent viscous damping, it is
first necessary to determine the substitute structure displacement (Eq,(5.3)) and effective
height (Eq.(5.5)). ‘The necessary calculations arc summarized in Table 5.1236 Priestley, Calvi and Kowalsky. Displacement. Based Seisinie Design of Structures
1
= 0.149 (14.99
2
Enya =0.05 + 0.565 (223
\
23m J
Option 3: hy = 500 mm; hz = 600 mm; M, = 0.6M3, From Eq.(5.19)
8, =0.5x 0.002475 “2-011 @,, =0.8x0.002475 7 0.0155
From Eq,(5.20b), the system yield displacement is:
2M.6,,+M,0,,
2M0,+MiB.r 4
1 20.6%0.011 14 + 0.0155
2M, +My
+29.4=0.386m (15.2in)
2x0.6+1
‘The system ductility is thus decreased to {4 = 0.609/0.386 = 1.58, and from Eq.(5
the equivalent viscous damping is:
En = 0.05 + 0.565 (! S81) (14Mp + 1.7My)/¢ G.45a)
Muex22 (1.0Mp + 1.0Mpe +Mp)x0.7 G.45b)
where Mg is the seismic moment. Note that Eq.(5.45b) incorporates the 30% reduction
for redistribution which would require a corresponding increase (in value, not percentage)
at other locations.
‘The recommendations provided in this text are that, in addition to the conventional
gravity load design (we again assume that Eq.(5.45a) applies, including the strength
reduction factor), the seismic design case is given by Eg.(5.45¢):
Mues = Me G.450)
Substituting the calculated values for dead and live load we obtain:
Mics: 2190.9 kNm (1689 kip in)
Myesz 2 (16.2415.2)X0.7 + 0.7 Mp = 64.1+0.7Mg kN (567+0.7Mg kip.in)
Mies = Me kNm
For conventional design permitting moment redistribution, the larger of the first two
values will apply. Thus for low values of seismic moment, the gravity load combination
will govern, For the approach suggested herein, the larger of the first and the third values
will apply.
Figure 5.14 plots design moment against seismic moment for the two approaches. For
low levels of seismic moment, Maes; governs for both approaches. It is seen that the
seismic moment starts to govern at a very similar level of seismic input with both
approaches, and that except for a very small range of seismic input, the approach
suggested herein is more conservative than the approach allowing moment redistribution,
This adds support to the simple recommendation that gravity loads be ignored in the
seismic load combination, and that design strength be based on the larger of gravity and
seismic moments.
Section 4.2.5 discusses material properties for intended plastic hinge locations, and it
is recommended that design values for seismic design exceed the specified, or
characteristic material strengths is accordance with fre = 1.33 re = 1lfp
As explained in Section 4.2.5, these values are recommended in recognition chat
material strengths will normally exceed specified strengths at the building age when
seismic attack occurs, and that use of artificially low material strengths does not reduce
the probability that inelastic response will occur at the design level of seismic input, nor
does it improve safety. The main consequence of design to artificially low material
strengths is an increase in cost, particularly for capacity-protected actions. In accordance253
Chapter 5. Frame By
300 7] Seismic
- edistributed
z
2
& 100 4
4
°
0 100 200 300
Seismic Moment (kNm)
Fig.5.14 Beam Design Moment Incorporating Seismic Loading Based on
Recommendations in this Text Compared with Conventional Design Permitting
Moment Redistribution of 30%.
with the recommendations of Section 4.2.5, strength enhancement of flexural
reinforcement may be considered when determining required flexural reinforcement
quantities, and the use of a Mlexural strength reduction factor for seismic actions of
intended plastic hinges is viewed to be inappropriate.
Methods for flexural design of beams are covered in all standard texts on reinforced
concrete design, and need no further elaboration here, though guidance relating to
material properties and use of moment-curvature analysis is given in the
recommendations of Section 4.2.6. Particular reference is made to [PI] for details on
contribution of slab reinforcement to flexural strength, and other matters specifically
related co seismic design of frames. It is noted, however, that the conventional
distribution of flexural reinforcement in beams, where negative-moment reinforcement is
concentrated near the top of the section, and posiive-moment reinforcement is
concentrated near the bottom of the section, as illustrated in Fig.5.15(@), can cause
problems with congestion and concrete placement, particularly in two-way frames, where
intersecting layers of reinforcement occur at the top of the joint region. If the same
quantity of flexural reinforcement is distributed down the sides of the section, as shown
in Tig.5.15(), the flexural strength of the beam will be essentially the same as for the
conventional distibution, as trial calculations will show, since a larger proportion of the
total reinforcement will act in tension for a given sense of moment, compensating for the
reduced lever arm from the centre of tension to the centre of compression in the section,
Distributing the flexural reinforcement down the sides clearly reduce
congestion254 Priestley, Calvi and Kowalsky, Displacement-Based Seismic Design of Structures
problems, and also improves bond conditions for beam reinforcement through the joint,
reduces joint shear stresses, and improves beam shear performancel™'
ee
/ eoc6e yf
|
(a) Conventional (b) Vertically Distributed
Fig.5.15 Alternative Distributions of Beam Flexural Reinforcement for
Reinforced Concrete Frames!!!
5.6.2. Column Flexural Design
As with beam design, design of columns is well covered in standard texts and needs
litde discussion here. However, a more conservative design approach than that for beams
is appropriate since the columns should be required to remain essentially clastic, except at
the column base, and possibly immediately below the roof beams. It should be noted that
column moments need to be amplified above values corresponding to the desiga lateral
seismic force to allow for:
© Potential overstrength capacity at beam plastic hinges resulting from material
strengths exceeding the values specified for design.
Dynamic amplification of column moments resulting from higher mode effects,
which are not considered in the structural analysis carried out in accordance with
Sections 5.5.1 or 5.5.2
* Biaxial effects for columns that form part of two-way frames.
Ir should also be noted that practice in the USA, requiring that the sum of column
moment capacity in a given direction framing into a joint should exceed 1.2 times the
sum of beam moment capacity in that direction, framing into the joint does not provide
adequate protection against columns forming accidental plastic hinges, with potentially
serious consequences.
(2) Beam Overstrength Capacity: The consequence of overstrength flexural capacity
peing developed at intended plastic hinge locations was discussed in Section 4.5. The
maximum feasible flexural strength of beam hinges can be assessed by moment-curvature
analysis, adopting high estimates of probable material strengths, and incorporating strain-
hardening of flexural reinforcement and confinement of core concrete, as outlined inChapter 5. Frame Buildings
Section 4.5, or conservative def
Section 4.5.2. Since the column is required to remain elastic, column dependable fles
strength should be required to satisfy Eq.(4.59).
ult factors may be adopted for simplicity, as discussed in
ral
‘These are discussed in some detail in Section 5.8.4.
(b) Higher Mode Effect:
(©) Biaxial Attack: When columns form part of a two-way seismic frame, biaxial input
must be considered, There will normally be equal probability that the maximum seismic
excitation will occur in any orientation (including diagonal) with respect to the principal
axes, and this will be accompanied by simultaneous excitation, normally at a lesser
intensity, in the orthogonal direction. Although the strength of a frame building in the
diagonal direction will be larger than in the direction parallel to one of the principal axes,
{see Section 5.7) it is probable that simultaneous development of beam plastic hinge
mechanisms will occur in both principal frame directions, unless the design displacement
ductility in the principal directions is very low. Consequently, since it is required that the
columns remain essentially elastic when beam-hinge mechanisms form, the columns must
be able to resist moments corresponding to simultaneous beam hinging in osthogonal
directions.
Fig.5.16 Plan View of Moment Input to Interior Column of a Two-way Frame
‘This is illustrated for the interior column of a two-way frame in Fig.5.16, where the
beam moment vector inputs are represented with double-headed arrows, using the right-
hand rule, All moments ate the values applicable at the joint centroid, and the suffixes P
and N indicate positive and negative beam moments respectively. The required sum of
column diagonal moment capacities of columns in the storey above and below the level
considered, measured at the joint centroid is given by256 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
YM 2V Moin + Moin) +(M gap + Maan) (5.46)
ar)
The most usual case for two-way frames will have the moment capacity of the beams
in the orthogonal directions being equal, and the columas having square cross-sections.
Experiments and analyses indicate that the moment capacity of symmetrically reinforced
square columns in the diagonal direction is very similar to the capacity in the principal
directions. Hence the required sum of column moment capacities in the principal
directions, above and below the joint centroid can be determined as
DMe = Y Mey =V2 (Moye + Maw) (6.466)
For an interior column, as depicted in Fig.5.16, the seismic axial force will be close to
zero, and hence the required column reinforcement can be determined based on the
assumption of probable gravity loads. For an exterior column, however, seismic beam
shears will generate significant axial forces, cither tensile or compressiv Fig.5.10,
eg) which should be considered when determining the required column flexural
(set
einforcement, It could be argued that permitting column hinging in the tension exterior
column is justified, since this in itself will not imply development of a soft-storey
mechanism, which would require all columns in a given frame to hinge, Although this
might appear technically and economically justified, it should be noted that exterior
columns are more susceptible to joint shear failure than are interior (cither one-way or
two-way) columns, and vertical joint shear resistance in these columns is normally
provided by under-utilized capacity of the flexural reinforcement in the columns, assured
by the capacity design procedure. IF the columns are permitted to hinge, then this
excess capacity will not be available for joint shear resistance, and special vertical joint
reinforcement will be needed.
Note that for diagonal attack the procedure for determining the required column
capacities, outlined in Section 5.5.2(b) needs sore modification. In Eqs.(.42) and (5.44)
the column shears Veqy and Vey2 must be replaced with the corresponding column shears
Veoo and Veprz developed under diagonal attack, For a structure with equal beam
strength and bay length in the orthogonal directions, Vepor = V2Vcor etc. The beam
moment input to the joint, given in Eq.(5.43) as Mp1, + Mpz,, must be replaced by the
vectorial summation of the beam input moments, given by the expression on the right
hand side of the inequality of Eq.(5.46(a)). It is also suggested that the base moment
capacity of the columns be increased to 0.7VcorH, where Veor is the ptincipal-direction
column shear.
It should be noted that the common force-based design approach of determining the
required column biaxial moment capacity from consideration of simultaneous seismic
input of 100% and 30% design intensity in orthogonal directions, using elastic modal
analysis, with the elastic moment demands then reduced by the code-specified force-
reduction factor (or behaviour factor, or ductility factor, as it is often known) will provideChapter 5. Frame Buildings 287
inadequate protection against formation of a soft-storey sway mechanism in the ground
floor columns, since the procedure implies input beam moments of only 30% of capacity
in one of the orthogonal directions. In fact, it is the displacement demand in the
orthogonal direction that should be reduced t0 30%, not the moment. If the design
ductility demand exceeds 3.3, then the full moment capacity should be expected to be
simultaneously developed in the (minot) principal direction, as well as the (major)
principal direction. For diagonal attack, 100%/30% is non-conservative for {>2.
Corner columns require special attention, since these can be subjected to high seismic
axial tensions or compressions from the response of both orthogonal frames incident in
che corner columns. ‘There is comparatively little experimental information on the shear
performance of corner columns when subjected to biaxial beam input and high variations
in seismic axial force from tension to compression. However, it should be noted that the
collapse of the newly-constructed Royal Palm Hotel in the 1993 Guam earthquake has
been attributed to failure of the (admittedly poorly reinforced) joints of the corner
columns, as shown in Fig. 5.17. A design option that should be considered for two-way
frames is to provide separate exterior columns for the orthogonal frames, rather than a
common corner column, as suggested in Fig. 5.18(b). Flexibility of the floor slab berween
the orthogonal exterior columns provides continuity without excessive shear transfer.
‘This has the merit of reducing the design moments, axial force variation, and joint shear
force on these corner columns. In some cases of low to moderate seismicity, the
architectural appearance of full two-way structural continuity into the corner column can
be achieved with an alternate design where the seismic frame consists of less bays than
the full number along the side of the building, ‘The seismic frame is connected to the non-
seismic columns, including the commer columns, by gravity beams without moment
connection to the columns. In New Zealand, frames have been constructed with reduced
beam depth adjacent to corner columns, and hence reduced moment input to the critical
columns, to produce a similar reduction in corner cohimn actions.
(d) Column-Base Flexural Design: Although it has been recommended above that the
seismic axial force be considered when determining the required column reinforcement
for capacity-protected exterior and corner columns, the same is not required when
designing the intended column-base plastic hinge. If the simplified analysis procedure of
Section 5.5.1 is adopted, then equal column-base design moments Mc would result for
the exterior tension and compression columns. If the axial force from the tension
column was used to determine the required amount of column flexural reinforcement,
then the moment capacity of the compression column would exceed the design moment
significantly. This is illustrated by the dashed line for the column axial-force/moment
interaction diagram shown in Fig.5.19. With this option, it is seen that the flexural
capacity of the tension column (with reduced axial force G-B) is matched to the required
capacity Mc. The capacity of the compression column, with axial force GFE is My. The
total resisting capacity of the structure, in terms of base overturning moment thus will
exceed the demand corresponding to the design distribution of lateral forces.258 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Fig.5.17 Failure of Royal Palm Hotel Corner Joints in the 1993 Guam Earthquake
(photo courtesy Gary Hart)
| =
(a) Conventional Layout (b) Alternative Layout
with Corner Column No Corner column
Fig.5.18 Plan View of Two-Way Frame Layout Options
An alternative approach would be to determine the required columa flexural
reinforcement based on the design moment and the axial gravity load G alone. This
option is represented by the solid line in the interaction di ‘The
reinforcement content Pz will be less than for the former alternative, and the momentChapter 5. Frame Buildings 259
capacities of the rension and compression columns will then be M; and M; respectively
Txamination of Fig.5.19 indicates that the average moment capacity will be very close to
he required value of Mc. Thus a mote efficient de
n has been achieved.
MM: Moment
Fig.5.19 Axial Force/Moment Interaction Diagram for Exterior Columns
Norte thar this could have been recognized at the start of the design process, where
che total base shear force was allocated between the columns, and the base moment
capacities were chosen (see Section 5.5.2(b)). A lower shear force and base moment could
have been allocated to the tension column, with a corresponding increase in the values
for the compression column, It will be noted that this is analytically a logical approach in
hat it recognizes that the tension column has lower elastic stiffness than the compression
column (see Section 4.4.2), and hence attracts less seismic shear and moment. It is
zecommended that 2 moment of 0.3 7h., where Tis the seismic tension force in the outer
column, and fie is the column depth, be subtracted from the tension column and added to
che compression column.
3.7 DIRECT DISPLACEMENT-BASED DESIGN OF FRAMES FOR
DIAGONAL EXCITATION 7
In the previous section the importance of designing columns of two-way frames for
diagonal excitation was discussed, A related aspect that requires consideration is the
control of drift when the excitation is in the diagonal direction. Implicit in the discussion
thus far
‘as been the
ssumption that design for specified drift or ductility limits in the
principal directions will provide a design that also satisfies drift or ductility limits in the
diagonal direction. We now show that to indeed be the case by considering a two-way
frame building with equal strength and stiffness in the orthogonal directions.
Consider the building response illustrated in Fig.5.20. Figure $.20(a) shows the yield
ign displacements in the principal (X and Y) and diagonal (D) directions in plan
view. Yield displacements are Ay and Ayy in X and Y directions respectively. For
and de260 Priestley, Calvi and Kowalsky, Displace
Based Seismic Design of Structures
in the two
simplicity we assume a square building with equal strengths and stiffnes
principal directions, so Ayy = Ayy = Ayp. If the building is displaced in the diagonal
Id will not occur until the displacements in the two principal directions are
equal to their yield values, and hence the diagonal yield displacement is
direction,
Ay, =V2A,) (5.47)
>
Bx Dia Ax Bed Aa An T) T»
(a) Design Displacements, (b) Force-Displacement (c) Displacement
Principal and Diagonal Response Spectra
Fig.5.20 Comparison of Displacement Response in Principal and Diagonal
Directions
“The relationship between design displacements in the principal and diagonal directions
will depend to some extent on whether strain or non-structural drift defines the design
limit. In the first case, the diagonal limit displacement corresponds to achieving the limit
strains in the principal directions, and hence the diagonal limit dis-
placement is Ap = V2Ajes. If the design displacement is limited by non-structural drift,
then it could be argued that the maximum displacement in the diagonal direction under
diagonal attack should not exceed the displacement limit applying in the principal
directions. This displacement is represented in Fig. by Apr where Apr = Aas
Clearly this is the more critical of the two cases, though it can be argued that a larger
design displacement is appropriate in the diagonal direction, even if non-structural drift
limits the design, since non-structural elements will normally be oriented parallel to one
of the principal directions, and achieving the design drift in the diagonal direction will not
correspond to critical non-structural displacernent.
If the design response in the diagonal direction exceeds the diagonal yield
displacement given by Eq.(5.47), then the strength of the frames will be developed in the
principal directions. Resolving the strength of all frames in the X and ¥ directions into
the diagonal direction, it is obvious that the diagonal base shear strength is V2Vsase whereChapter 5. Frame Buildings 261
View is the design strength in each of the principal directions.
“The force-displacement response of the structure in principal (P) and diagonal (D)
directions is represented in Fig.5.20(b), where it is assumed thar the diagonal
clisplacement response is either Ap) or Apz. As explained above, we consider only the
more critical case of Apy = Ayes. We first assume that the structure is designed in the
principal directions, and then checked to determine whether or not response in the
diagonal direction satisfies the displacement limit, It is seen thar if the structure achieves
the design displacement in the diagonal direction, the stiffness Kep is greater than the
stiffness Kep in the principal directions. This would appear to imply that the displacement
demand in the diagonal direction will be less than in the principal direction, since the
effective period will be higher (Bq.5.10). However, the ductility demand in the diagonal
direction will also be less than in the principal direction, since the yield displacement is
larger (Eq.5.47) and hence the effective damping will be less, and displacement will
increase. The two conditions — principal and diagonal — are represented in the
displacement spectra of Fig. 5.20(c). Since the effects of reduced period and reduced
damping are counteractive, it is not immediately clear whether displacement response will
be larger or smaller in the diagonal direction.
For a given level of ductility, #/in the principal direction, this can be resolved by an
iterative analysis ¢s follows, making the initial assumption that the diagonal response
displacement is equal to the principal response displacement. Diagonal and principal
stifinesses will thus be related by
Ky= V2-Kp (5.48)
Since the effective period is given by T,=2rty/m,/K, , and the mass is the same
whether principal or diagonal attack is considered, the diagonal and principal effective
periods will be related by
Ty = Tp M2 =
8417, 6.49)
Diagonal and principal direction ductilites are related by
Up =Mp V2 6.50)
Assuming a reinforced concrete building, the ratio of effective damping in diagonal
and principal directions can be found from Eq,(5.9a) as
0.05+0.565(U,
0.05-+0.565(u,
Wi pt
Vey 4
(5.52)
Wi up 652)264 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Moment (No Units)
(a) Beam Moments
E+ Ge ——_»
Bt Ge 2
10
Shear (No Units)
L
+
(b) Beam Shears
Fig.5.22, Design Moments and Shears for Beams
forces, and corresponding overstrength seismic moments are Z and E? respectively. In
accordance with the recommendations of Section 5.6.1, flexural design of the beam
plastic hinges is based on the larger of FG and E, and it is seen that the seismi
governs, For the regions between the beam plastic hinges, design moments are found
from the combination of reduced gravity loads applicable for the seismic design
combination, and overstrength moment capacity at the beam hinges. Since the beam
moments cannot exceed the overstrength values at the plastic hinge locations, the design
case
moments within the beam span Zg are found by adding the gravity momentsChapter 5. Frame Buildings 267
Note that for two-way frames, @ is applied to the columa moment Me resulting from
design forces corresponding to one-way action. Note also chat ay is height dependent,
and equal to 1,0 of 1.1 at base and top of the column for one-way and two-way frames
respectively, since column base hinging is expected at the base, and permitted at the top.
It will be noted that with the minimum value of g = 1.47 appropriate for the New
Zealand code, the minimum required column moment capacity will be 2.48 or 2.64 times
the moment corresponding to lateral design forces, for one-way and two-way designs
respectively. It should also be noted that the moment amplification is independent of the
level of design ductilicy demand.
One-way Frame Oy Toney Frame iy
Fig.5.23 Column Dynamic Moment Amplification Factor for Flexure in N.Z.
Design'?!
In Eurocode EC8M higher mode effects for column moments are directly considered
by combination of the modal moments by either SRSS or CQC combination rules.
(b) Results of Inelastic Time-History Analyses: Pewtinga and Priestley! report
results of dynamic amplification of column moments in one-way frames of 2-storeys (0
in height, designed by DDBD principles for displacement ductilities of about
C8 clastic modal superposition approach was
20-storey
2.7. These results indicated that the
generally non-conservative, while the New Zealand approach was generally significantly
over-conservative. The analyses were carried out for different levels of seismic intensity,
corresponding 10 50%, 100% and 200% of the design intensity. It was found that
although the base-shear strength of the buildings remained the same as the intensity was
changed, the dynamic amplification of the column moments increased with intensity,268 Priestley, Calvi and Kowalsky. ‘Displacement-Based Seismic Design of Structures
except at the column base, where moment was controlled by the moment capacity.
Results for four-storey to sixteen-storey frames are shown in Fig.
In Fig.5.24, the column moments are the sum of all column moments in the different
columns at a given level, and the absolute maximum values, without sign, have been
plotted, Lines TH I = 0.5, 1.0 and 2.0 represent the average results from five spectrum:
compatible accelerograms at intensities of 50%, 100% and 200% of the design intensity.
Note that the DDBD design moments for she columns, shown by dashed lines in
Fig.5.24, were calculated in accordance with the suggestion of Section 5.5.2(b) that the
input moment-sum from beam plastic hinging be equally divided between the columns
above and below the joint. The analyses do not include material overstrength, except for
the component resulting from strain-hardening, which is a smal] component of the total.
Strain-hardening of column-base reinforcement is responsible for the small amount of
moment increase with insensity, at the column base. Note that in the lower third of the
building height, column moments at the top of each storey are significantly higher than at
the bottom of the same storey, and that the difference increases as the seismic intensity
increases, This is partly a result of anchoring the column-base moment in the bottom
storey at the plastic hinge capacity. It is clear from these resuks that dynamic
amplification of column moments is dependent on the ductility demand, since this (for a
given design) increases approximately proportionately with the intensity of excitation. Ie is
also apparent from examination of the plots that the amplification factor of 1,6 suggested
in [P1] for the central part of the column height is excessive at the design intensity, but
quite reasonable at 200% of the design intensity
(©) Design Recommendations: The data in Fig,5.24 lead to the following revised
approach for determining the required column flexural strength:
6M, 20°O/M,
In Eq(5.60) the overstrength factor @? is calculated in accordance with the
recommendations of Section 4.5
© The dynamic amplification factor @ is height and ductility dependent as shown.
in Fig.5.25, where from the first storey to the ¥ point of structure height
-15+0.13(u° =1) 6.612)
=1.00 6.61b)Chapter 5. Frame Buildings
269
4 Storey Column Moment Sum
Envelopes
ay ]
1
Level
1 | me Desi
maisos
THI 20
ol
° 5000 10000 S000
Moment (km)
8 Storey Column Moment Sum
Envelopes
ay pe
a Ae bee
THI=05
© 5000 10000 15000 20000 25000
Moment (kNm)
0
°
12 Storey Column Moment Sum
Envelopes
THis9s
ee tai=20
x a
Moment (kNm)
Fig.5.24 Column Moment Envelopes from Time-History Analyses at Different
Seismic Intensities, Compared with Design Envelopes!"Chapter 5. Frame Buildings 2
conservative to ensure that no incipient plastic hinging can develop in columns. tt will be
recognized, however, that the suggested approach is more conservative than all except
existing New Zealand provisions for column design
5.8.5 Column Shear
(@) Existing Methods Accounting for Dynamic Amplification: Seismic design codes
generally include more restrictive provisions for the capacity-design shear forces. in
columns, in recognition of the potentially catastrophic consequences of column shear
failure. The US. IBC codetl is perhaps the least conservative, as its provisions are
related only to flexural overstrength issues. Dynamic amplification of column shears is
not directly addressed, The basic requirement is that the design shear force shall be at
least as high as the lower of the shear corresponding to overstrength beam moment input
(based on a yield strength of 1.25f, and no flexural strength reduetion factor), and the
shear corresponding to development of probable moment capacity (at 1.23f,) at the top
and bottom of the column.
New Zealand practice again follows recommendations in [P1], which were based on
earlier time-history analyses. Apart from in the first-storey columns, the maximum
column shear is amplified for overstrength and higher-mode effects in accordance with
F.g.(3.55), where
For one-way frames, @, 5.638)
(6.63b)
For two-way frames, Qs
For first storey columns, hinging is expected at the base, and may also occur below the
first-floor beams as a result of beam exterision!®!, even when protected by the design
approach suggested in Section 5.5.2(b). Consequently the column shear in the critical
first floor column is calculated assuming overstrength hinging at both locations:
Pay =i Me
(5.64)
where He is the clear column height to the soffit of the Level 1 beam,
Eurocode EC8! specifies that column shear force shall be determined by modal
superposition, Problems with modal superposition for shear forces are discussed in some
detail in Chapters 6 and 7. It is sufficient co note here that elastic modal superposition is
non-conservative for ductile structures because it incorrectly assumes that the higher
mode shears can be reduced by the ductility, or behaviour factor applicable to the first
mode response.
(b) Results of Inelastic Time-History Analyses: The analyses reported in [P17] and
described above, also provide information on dynamic amplification of column shear.272 Priestley, Caivi and Kowalsky. Displacement-Based Seismic Design of Structures
Results of the analyses at levels of seismic intensity between 50% and 200% of design
intensity are shown in Fig.5.26 for four-storey to sixteen-storey buildings. Results for
owo-storey and twenty-storey followed similar trends and are reported in [P17]. As with
the column moments, the design shear envelope corresponding to the DDBD lateral
forces is included as a dashed line, and values plotted correspond to the sum of all
columns in the frame at a given level. Material overstrength was not modelled in the
analyses, except for the fraction resulting from strain-hardening of reinforcement
Results follow similar tends to, and are compatible with those noted for columa
moments, in that column shear force increases with increasing seismic intensity. As
discussed above, this can be interpreted to imply that dynamic shear amplification
increases with ductility, since ductility increases with intensity for a given structure. ‘This
influence is not included in existing design methods. It will also be seen that the shape of
the time-history envelopes is generally similar to the design-force envelope, though a
constant offset of shear demand above design-force envelope with height would seem
more appropriate, particularly for the lower structures, than a constant multiplier as,
recommended in [P1]
(c) Design Recommendations: Based on the above observations, the following form
of the dynamic amplification factor for column shear is propose
‘ M?+Me nom
Oy 2OV 5 +0. ctue SG 6.65)
‘The following comments relate to Eq.(5.65):
* The overstrength factor @ is calculated in accordance with the recommendations
of Section 4.2.5.
* The shear demands Vg correspond to the design
from the DDBD process. Vezsase is the value of Ve at the base of the column,
The system ductility {1 is not reduced by ¥ since this factor cancels with the
same amplification of Ve.
eral force distribution found
© The upper limit of column shear corresponds to development of plastic hingi
at overstrength capacity at the top and bottom of the column,
clear column height, Ho
* The approach suggested provides a good estimate of the shear demand up the
height of two-storey to eight-storey frames. For taller frames, the shear force at
roof level predicted by Eq.(5.65) becomes increasingly conservative, but is likely
to be controlled by the upper limit. Generally, prescriptive confinement
requirements will govern transverse reinforcement design in the upper levels
As discussed in Section 5.6.2(0), and 5.8.5(a), maximum feasible column shears will be
increased in two-way frames under biaxial attack. The shear force may be found from the
vectorial addition of the principal direction shears. The following considerations appiy:
ngs
eparated by theChapter 3. Frame Buildings 273
4 Storey Shear Envelopes 12 Storey Shear Envelopes
a 1 =P Py
Rly Zn ae Dain
! 0 timos
\ ee tiinso
3 - °
i ‘ U ——miie20
! ;
'
2 - ‘
i
' 5
i 4
ypc :
rates
2
ee}
i | ‘|
eH 20
ol +. L o |
200040006000 g000 suo 1000015000
Shear (kN)
8 Storey Shear Envelopes
© 3000 6000 900012000
Shear (kN)
Shear (kN)
16 Storey Shear Envelopes
‘i
THO
Hie
30006000 990912000
Shear (kN)
Fig.5.26 Column Shear Envelopes from Time-History Analyses for Different
Seismic Intensities Compared With DDBD Envelope!"274 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
© Shear strength of square columns is essentially the same in principal and
diagonal directions,
Maximum diagonal ductility demand is less than the principal direction
demand by a factor of V2 (see Section 5.7)
Thus, for square columns of ewo-way frames, the required nominal shear strength in a
principal direction, is given by
Mo sMe
V29°V,, +0.14LY, og MMe (5.66)
Vy te Hl,
where the suffices » and 2 relate to actions for one-way and two-way response
respectively.
5.9 DESIGN VERIFICATION
Much of the design recommendations in the previous section have resulted from an
extensive analytical study described in derail elsewhere”, though the dynamic
amplification factors for shear and moment are new. In this section we briefly compare
design predictions with the average results from time-history analyses reported in (P17].
5.9.1 Displacement Response
‘The full analytical stady investigated 2, 4, 8, 12, 16 and 20 storey regular frames. We
limit comparison to 4, 8, 12 and 16 storeys, in the interests of brevity. Results for the 2-
storey were similar t0 the 4-storey, and results for the 20-storey were similar to the 16-
storey frame,
Figure 5.27 compares the average time-history response for an intensity ratio of I=1.0
with the design profile in terms of displacement envelopes. Also shown is the average
time-history response for intensity ratios of I=0.5 and 2.0, for comparative purposes. In
all cases the drift envelopes (see {P?7)) had peak valves within 10% of the design limit of
0.02. For the shorter frames the maximum drift was not greatly affected by higher mode
effects, and peak drifts occurred at the first storey, For the 16 and 20-storey frames,
higher mode drift amplification was substantial, and maximum drifis occurred at about
80% of the frame heights
Comparison of the design displacement profile and the
indicates excellent agreement in all cases,
=1.0 time history results
5.9.2, Column Moments
Figure 5.28 compares the sum of column moments at a level with the predictions
based on dynamic amplification in accordance with Eq,(5.61) and Fig.5.25. Note that the
time-history analyses were carried out using the design expected values for the materialChapter 5, Frame Buildings
4 Storey Displacement Envelopes
r
\ ~
THI-08
eee
ei
0.00 0.10 920 0.30 040 050
placement (a)
8 Storey Displacement Envelopes
Lave
Displacement (n
060
12 Storey Displacement Ei
‘elopes
Level
6
‘
‘|
a] em bai
3 Tal=o5
2 ete r0
ee i=20
boa 02s 080 075 100 125 1.50
Displacement (m)
16 Storey Displacement Envelopes
16 t
1
7
‘
s
7 Ae Dees
3 mHi-os
2 isto
°
0.00 025 050 075 100 125 180 1.78
Displacement (m)
Fig.5.27 Design Displacement Profiles Compared with Average Time-History276 Priestley, Calvi and Kowalsky.Displacement-Based Seismic Design of Steuctures
Results!P""1
Height Suories),
0 4000 080 12000 © 40008000 12000 0 4000 g000 12000,
Momene (kN) Moment («Nm) Moment (ki
(a) # Storey, F (b) 4 S:orey, I=L0 (©) 4Storey,
© 4000 8000 12000 16000 0 80006000» 91000020000
Moment (kNm) ‘Moment (km) ‘Moment (kNim)
(@) 12Storey, 105 (©) 12Storey (8 128rorey, 1=2.0
Height (Stories)
° a
TTT
© 4000 8000 12000 i600 0 800016000 = 01000020000
Moment (kNen) ‘Moment (kNm) Moment (Nm)
(g) 20-Stomey, I=0.5 (@ 20-Storey, 1=2.0
Fig.5.28 Capacity Storey-Moment Envelopes Compared with Time-History
Results (D = DDBD; TH= time history results; CD = Eq.(5.61))Chapter 5. Frame Buildings 27
strengths, and hence the material overstrength factor was taken as @’ = 1.0. In fact some
strain hardening beyond that accounted for in design occurred for intensity ratio I=2, and
aslightly higher amplification of the design moments would be appropriate.
Note that the moments plotted in Fig.5.28 are the sum of the moments in all columns
aca given level, and ate plotted without sign. The design moment envelope is identified
by “D®. Generally peak moments at one end of a column do not occur at the same time
instant as peak moments at the other end of the column, It is seen that the simple
modifier suggested in Eq.(5.61) provides a good envelope for all the three frames
depicted, at all three levels of intensity (and hence of ductility). It will be noted that in a
few cases maximum columa moments {rom the time-history analyses exceed the capacity
design envelope (identified by CD in the figures) by a small amount. ‘This should ot be
of concern, since the resulting ductiiry demand on the columas will be very low. It is felt
chat providing an absolute envelope to the time-history results is not economically
justified, given the substantial ductility capacity of modern well-confined columns,
5.9.3. Column Shears
Column storey-shears (ie. the sum of shears in all columns in a given storey) fom
time-history analyses at different seismic intensities are compared in Fig.5.29 with the
shear envelopes defined by Fq,(5.65). Note that the upper limit provided in Eq..65) by
the shear associated with column hinge formation at top and bottom of a columa has not
been considered in the time-history analysis, nor in the capacity design envelope.
As with the column storey-moments, the capacity design shears envelope the time-
history results except in a very few cases. The degree of conservatism is not high, though
it could pethaps be considered excessive for the I = 2, 20-storey case in the upper half of
the frame, At the design intensity level (1 = 1.0) the agreement between the time history
and capacity design envelopes given by the simple design approach is very satisfactory.
An alternative formulation of capacity design effects for columa moments and shears
which involves specific consideration of higher mode contributions has been proposed in
[P17]. This approach is more cumbersome 10 use, and does not give significantly
improved agreement, If more accurate estimation of capacity design column moments
and shears is required, it is recommended that inelastic time-history analysis of the
structure be carried, based on the principles outlined in Section 4.9.
5.9.4 Column Axial Forces
In determining the reinforcement requirements for flexure and shear in the columns,
the axial forces also need 10 be known. These can be found from the gravity loads, plus
the axial forces contributed by the beam seismic shear forces. Note that for corner
columns, the beam shears from both orthogonal directions must be considered,
potentially doubling the seismic contribution t column axial force. If the resulting axial
forces, particularly the uplift force on the corner tension columa are too high, redesign
may be appropriate, as suggested in Section 5.6.2(¢), to reduce the cokima axial force278
Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
© 2090 4000 6000
‘Shear (kNen)
(8) 4Storey, 1-05
Height (Stories)
0. 4000
Storey Shear (KIN)
8000
(8) 12-Storey, 120.5
Height (tories)
0 _ 4000, 8000
Storey Shene (kN)
(g) M-Storey I-05
8009 0
120000
Storey Shear (KN)
(b) 4 Storey I=L0
to
40008000 12000
Storey Shear (KN)
(©) Storey T=10
20
16
2
4000, 8000
Storey Shear (KN)
0
000
(1) 20-Storey
2000 4000 6000 8000. 0
opt
12000 0
2000 4000 6900 8000
Storey Shear (KN)
(c) 4Storey, 1=20
—
©. 40008000
Storey Shear (kN)
() RSe0rey 1-29
2000
40008000 12000
‘Storey Shear (KN)
() 20-Storey 1=2.0
Fig.5.29 Capacity Storey-Shear Envelopes Compared with Time-History Results
(D = DDBD; TH = time history results; CD = Eq.(5.65))280 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Design Base Shear Force: The steps to determine the frame design base shear force are
listed below. Relevant data from Table 5.1 are included io the first three columas of Table
53.
From Bq.(5.4) the substitute stracture equivalent mass is (refer Table 5.3,Col.(3))
Dlm,A,)/4, =371.5/0.609 = 610 tonnes (1345 kips)
The reduction factor to be applied to the 5% damping displacement spectrum for
13.1% damping is given by E.q.(2.8) as
0.07
0.02+0.131
os
. ) = 0.680
‘Thus the corner displacement at T = 5.5 see for € = 13.1% damping is 0.680x1.4
=0.953m, The corresponding displacement spectrum is shown in Fig.5.30 as a dotted
line. By proportion, the effective period of the substitute structure is
0.609
T, = 55-22 =3.Sisec
0.953
Table 5.3 Calculations for Example 5.2 (1kN = 0.225 kips, 1 m =39.4 in)
Tevel,i ] Height | miA; F Vs OTM | V.beam! | M.beami*
Hi (m) (N) | kN) | Nm | (kN) | (Nm)
@) ® (4) ) (10) (it) (12) (13)
R 43.0 57.95 | 286.1 | 286.1 i) 694 138.8
tt 305 4686 1351 | 4213 | 1001.4 [1024 2043;
10 36.0 43.84 126.4 547.6 2475.7 132.8 265.6
9 325 4059 71 | 6646 [43923 [161.1 322.3,
8 29.0 37.13, 1071 | 7717] e786 | 187.1 3742.
7 255 33.45) 96.5 8682 | 94196 | 2105 421.0.
6 22.0 29.55) 85.2 953.4_| 124582 | 251.2 462.3
5 185 25.43 73.3 | 10267 | 157949 | 248.9 497.9
4 15.0 21.09 60.8 1087.5 | 193882 | 263.7 527.3
3 5 16.53 477 | 11381 | 231004 | 2753 550.4
2 80 175 339 | 1169.0 | 271672 | 2834 566.8
T a5 731 211 1190.1 | 312586 | 288.5 S77
o 0 0 00 1190.1 | 366139 | 0.0 0.0
Sum 37147 1415.0 | 101211 2454.0
* At column faceChapter 5. Frame Buildings 281
‘The substitute structure stiffiness is found from Eq.(5.10)
K, =4x°m, | 7m? x610/3.51? =1955KN /m
Hence, from Ei
(5.11), the design base shear force is
F Voge = KA, =1955X0.609 = 190KN (16.5% of weight).
Distribution of Base shear to Floor Levels: The modified form of Eq,(5.12) allowing
6.27) is used to determine the floor
for an additional force at roof level, given in
forces:
‘base
F=(F)+0.V gue (MA,)/ DmA, where F,=0.1,,,. at the roof, and
zero elsewhere, Hence
F, = (119) +1071x(m,A,)/371.5
Values for Fare tabulated in Col.(9) of Table 5.3.
Storey Shear Force and Overturning Momen found by
summing the floor forces above the storey considered. These are listed in Col.(10) of
Table 5.3. Storey overturning moments at the floor levels are also found from the lateral
Storey shear forces ar
fore
OTM,
PF (4,-1,)
‘The overturning moments at each Jevel are tabulated as Col.(11) in Table 5.3. The base
overturning moment for the frame is OTMpme = 36,614 kNm.
Column Base Moments: \n Section 3.5.1 it was recommended that the point of
contraflexure in the ground floor columns be set at 0.64% for one-way frames. For two-
way frames, Section 5.6.2(c) suggested this be modified to 0.7% In this design we adopt
a compromise, at 0.654%. Note that we use this value for determining one-way response
The actual point of contraflexure for biaxial response will be lower.
‘The tonal eesisting moment provided at the column base will thus be
YM, =V pne X 0.65, =1190% 0,654.5 = 3481kNer
This is less than 10% of the total base OTM.286 Priestley, Calvi and Ko
ky. Displacement-Based Seismic Design of Structures
research in precast frame construction has concentrated on prestressed beams utilizing
unbonded rendons!?!%", Although prestressed concrete has not generally been favour
m-column sub-
for seismic resistance, tests on both i sitd®\ and precast}
assemblages with fully grouted tendons have developed ductility comparable to
monolithic reinforced concrete. However, after moderate ductility levels had been
achieved, these subassemblages suffered excessive stiffness degradation at low
displacements. This reduction was caused by a reduction in effective prestress clamping,
force through the column, resulting from inelastic strain of the prestressing tendon at the
ctitical section, This behaviour is shown in idealized form in Fig. 5.31(a) which refers to
8 typical prestressing steel stress-strain curve,
150
3 50
2 is
F so
3 aso
450
750
273 46S O85 OSS 168 298
Displacement (in)
(a) Prestress loss due to inelastic response (b) Force-displacement hysteretic response
Fig.5.31 Seismic Response of a Beam-to-Column Unit Prestressed with Bonded
Tendonst¢#l
In Fig5.31(2), fi is the initial steel stress after prestress losses. During low-level
seismic response, fluctuations of the steel stress will be within the elastic range, and no
loss of prestress will result. At a ductility level of (say) ff =2, represented by point 2 in
Fig.5.31(a), the maximum prestressing, steel response is expected to be on the inelastic
strain curve. On unloading, the steel follows a linear descending
© the initial elastic portion. Hence, when the structure returns
portion of the stress
branch essentially parallel
to zero deformation, the effective steel stress is reduced to fy
On unloading from higher ductility levels, involving larger inelastic strains as
indicated by point 3 in Fig,5.31(a), the entire prestr ay be lost. This is clearly
undesirable, particularly for precast connections where the surfaces beoween the beam
and column are planar, rather than a rough naturally forming crack as may be the case for
monolithic reinforced concrete construction, Prestress shear-friction, which would be
hear forces from the heam to the column would then be
relied on to transmit gravity
ineffectiveChapter 5. Frame Buildings 287
The result of this behaviour is excessive pinching of the force-deflection hysteresis
loops, as indicated in the typical response of a prestressed subassemblage, shown in
Fig,5.31(bJ4, Response at low ducclity levels is satisfactory, and peak lateral strength is
maintained at high displacement ductilities. After displacing to these high levels, however,
laceral stiffness is almost completely lost at low displacement levels. ‘The test results
shown in Fig.5.31(b) refer to a test unit without additional gravity load on the beams. If
the gravity loads been modelled in this test, the performance would have been even less
desirable.
5.11.2 Prestressed Frames with Unbonded Tendons
(a) Performance Advantages: ‘The use of unbooded, or partially unbonded tendons,
where the tendon is unbonded through the joint and for some distance on either side, as
indicated in Fig.5.32(2) has been suggested as a means for improving performance!?24,
‘The following advantages were proposed:
#1 the length of debonding is sufficient, the required ultimate displacement could
be achieved without exceeding the limit of proportionality of the prestressing
steel. Consequently there would be no loss of prestress on unloading from the
design level of displacement. Shear friction on the beam-to-column interfaces
would be maintained at all response levels, and support of gravity loads would
not be jeopardized.
* Design of the beam-to-column joint region would be simplified, since the joiat
shear forces would largely be transferred by a diagonal strat, as shown in
Fig.5.32(b). This is a consequence of the prestress forces on either side of the
joint being equal for each tendon, because the tendons are debonded through the
joint. ‘Thus, reduced levels of special joint reinforcement would be needed.
© The response would be essentially elastic, though nonlinear, as indicated in Fig.
5.32(©). Although it is recognized that this would have possibly undesirable
consequences for energy absoxption, it has the merit that, following response to
the design level earthquake, the structure would return to its original position
without residual displacement, and the initial stiffness would be restored.
() Lateral Force-Displacement Characteristics: Force-displacement characteristics
of a typical debonded precast beam-to-column prestressed concrete subassemblage are
indicated in Fig.5.33(b), based on the forces and relative displacements measured at the
cop of the beam-to-column subassemblage, as shown in Fig,5.33(a)-
‘Assuming no tension capacity across the beam-to-column interface, nonlinearity of
response will initiate at point 1 (Fig.5.33%)), when pre-compression at the extreme fibre
is lost, and a crack starts (© propagate, Assuming further that the prestress centroid is at
mid-height, the corresponding moment is
M,,=Th,/6 (6.67)288
1ed Seismic Design of Structures
e
Debonded
Length
t | £
- L|
ney > 1
LE x
Prestess
‘Tendons
(©) Idealized force-deformation characteristic
Fig.5.32 Beam-to-Column Connection with Debonded Tendons
for a rectangular section, where T; is the initial total prestress force, and Ay is the beam
depth. For the dimensions shown in Fig.5.32(a), the corresponding lateral force Fwill be:
(6.68)
where he is the column depth and Le is the column height between contraflexure points
The corresponding displacement can be found from simple linear elastic analysis based
‘on uncracked section properties.Chapter 5. Frame Buildings 289
1
'
t
1
_f
Le
A
(a) Subassemblage Configuration (b) Non-linear Elastic Response
Fig.5.33 Force-Displacement Response for a Debonded Prestressed
Beam-to-Column Subassemblage
When the crack has propagated to the centroidal axis, corresponding to point 2 in
Fig.3.33(), the moment will be:
M, =Th,/3 6.69)
Deviation from the initial elastic stif
unless the average prestress level is fairly high fe > 0.25f"2)
Beyond point 2 the precise force-displacement relationship is tess simple to determine,
since steel and concrete strains are not linearly related, However, since the aim will be to
develop a bilinear approximation to the force-displacement response (sce Fig. 5.32()), it
is sufficient to compute the conditions at maximum displacement. In terms of design
requirements, this may be taken as cortesponding to the limit of proportionality of the
prestressing steel. At this point, the tendon force, and also the tendon elastic extension
above the initial condition are known. Assuming confined conditions for the concrete in
fess between points 1 and 2 will be minimal,
the plastic hinge region the corresponding moment capacity Mj can thus be found from
standard section analysis, Further, assuming that the opening of the crack at the beam
column interface is directly related to the tendon extension by simple geometry, the
displacement at point 3 can be shown!” to be:
(5.70)
where fp is the steel stress at the limit of proportionality, x is the unbonded length of
tendon on one side of the joint (see Fig. 5.32(a)), and cis the depth from the extreme
compression fibre to the neutral axis in the beam plastic hinge region (found from the290 Priestley, Calvi and Kowalsky. Displacement-Bosed Seismic Design of Structures
known confined concrete stress-strain parameters). Note that in all designs and
experiments to date using unbonded tendons, the unbonded length x has been taken as
half the beam span length. ‘That is, the tendon has been fully debonded along its length,
except at the anchorage. Moze complete details are provided in [P20]
Large-scale testing of interior and exterior beam-to-column subassemblages!?") has
shown the expected benefits of low residual displacement, high retained elastic stiffness,
and reduced need for transverse reinforcement, They have also shown thar damage levels
are very much less than are obtained with equivalent reinforced concrete structures
Figure 5.34 shows condition of an interior joint at the design drift of 2)
joint at 4%, twice the design drift. In both cases only superficial damage had occurred.
Cracks in the joint region fully closed when lateral force was removed,
fo, and an exterior
(a) Interior Joint at 2% drift (b) Exterior Joint at 4% drift
Fig.5.34 Damage to Beam-to-Column Precast Units with Unbonded
Tendons!"1
5.11.3 Hybrid Precast Beams
(4) Conceptual Issues: \ disadvantage with beams using unbonded tendons is thac
cis litle additional damping provided by the inelastic response, and the response is,Chapter 8, Frame Buildings 291
as noted above essentially non-linear elastic. ‘The implication is that displacements will be
larger than with equivalent reinforced concrete structures, or, conversely, that the DDBD
process will result in a higher required design force for a specified drift limit. Behaviour
can, however, be improved by adopting a hybrid design, as illustrated in Fig.5.35.
f L$ oun cowarvonae aenronene
280% uneowoen post—rensoneD
SoS Wve REDE WW
2
‘\
rf
NT W/ rae O80
a 10 STRESSN.
Fig.5.35 Hybrid Beam-to-Column Connection used in the UCSD 5-Storey
Precast Building Test!
This detail, which was one of several connection details tested in the 5-storey precast
building test!S!”l carried our at the University of California, San Diego in 1999 uses a
combination of central unbonded post-tensioning and conventional mild-steel
rcinforcement grouted into corrugated ducts in the beams and column at the connection.
‘The mild-steel reinforcement yields sequentially in tension and compression under eyelic
loading, dissipating energy, and adding to the flexural strength of the connection, but the
clamping force of the unbonded tendon is sufficient to yield the reinforcement in
compression when the lateral load is removed, hence removing any residual displacement.
The resulting hysteresis loop is flag-shaped (see Section 4.9.2(8) and Fig.5.36)
Figure 5.36 illustrates the moment-rotation characteristics for a typical hybrid design,
with a central prestressing tendon of area Ap, and mild steel reinforcement areas of As at
top and bottom of the section, separated by a distance d’ (see Fig.5.36(a)).
‘The response of the section based purely on the prestressing force is depicted in
Fig5.36(b) by the solid line, The moment capacity Mp is found as described above, based
‘on conventional section analysis, Since the elastic portion of the moment-rotation
response is governed by the uncracked stiffness of the beam section, the addition of mild292 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
steel reinforcement has negligible effect on the initial elastic branch of the response, but
the moment capacity is increased by the reinforcement couple Ms = Asha? as a
consequence of the reinforcement yielding in tension at the top, and compression at the
bottom of the section (or vice versa, depending on the sign of the moment). The total
moment-rotation response is depicted by the upper dashed line in Fig.5.36. When the
loading direction is reversed, the yielded reinforcement must yield in the opposite sense
(ie. tension to compression, and compression t0 tension) before open cracks can close.
‘The response on unloading is thus given by the lower dashed line on the tight side of Fig
5.36(b). To ensure that the advantages of low residual displacement are obtained, it is
necessary that Ms < Mp.
loment ©
Moment TM
’
a---7
(a) Beam Reinforcement (6) Moment-Rotacion Relationship
Fig. 5.36 Moment-Rotation Response of a Hybrid Beam-Coiumn Connection
(b) Yield Displacement of Hybrid Frames: "The increased stiffness of the prestressed
beams, compared with equivalent conventionally reinforced sections reduces the yield
drift, and hence increases the displacement ductility capacity of hybrid frames, for a given
design drift. Beam flexural and shear displacements, and joint shear deformation will all
be significantly less, but column flexural and shear deformations will be unaffected. Based
on these considerations, yield drifts are estimated to be about 40-50% of those for a
conventional concrete frame seinforced with mild steel of yield strength 400 MPa,
Conservatively using the upper limit, the yield drift can thus be writven as
8, = 0.0005 671)
h,
Since hybrid designs will normally be incorporated in building designs as perimeter
frames, with interior frames designed for gravity support only, beam span-to-depth ratiosChapter 5. Frame Buildings 293
will tend to be less than for two-way reinforced concrete frames, and yield drifts will be
further reduced. In conventional force-based design this would imply increased design
forces as a result of the reduced natural period, In DDBD, the reduced yield
displacement results in increased ductility demand, and hence increased damping. As a
consequence, the design force level is reduced.
(©) Equivalent Viscous Damping of Hybrid Frames: Example 3.2 (Section 3.4.3(d))
presents a method for determining the equivalent viscous damping for a flag-shaped
hysteresis rule where the height of the flag is P= 0.75xB,, (BF; is the yield force) and the
post-yield stiffness is r= 0. This can be generalized for other values of Band f, giving the
hysteretic component of the arca-based viscous damping as
- 4A Blu!)
2aF A, ur(l+r(u-))
Sou
2
8
Note that in relation to the formulation given in the previous section (Fig.5.36) the
flag height may be expressed as
2M.
p= 6.73)
M,+M,
‘The damping calculated from Eq.(5.72) is multiplied by the correction factor from
Fig.3.15, and added to the elastic damping component from Table 3.2 or Fig.3.13
5.11.4 Design Example 5.3: DDBD of a Hybrid Prestr:
including P-A Effects.
d Frame Building
Figure 5.37 shows details of a six-storey frame building of 25 x 25m (82 x 82{t) plan.
The seismic bracing system consists of peripheral frames of precast hybrid beams and
columns with internal non-seismic frames designed to carry gravity loads, and with
sufficient displacement capacity to tolerate the seismic design drifts. As suggested in
Fig.5.18(b), problems with corner columns, which are exacerbated in frames containing
P sed beams because of the need for intersecting anchorages, are avoided by
terminating the beams before the corners, and linking to the orthogonal peripheral frame
with either slabs, or non-seismic beams. Dimensions, material properties and floor
weights are included in Fig.5.37
The building is to be designed for a maximum drift of 0.025 when subjected to the
design intensity represented by the displacement spectrum of Fig.5.38, and P-A effects
are to be considered in the design.294 Priestley,
Displacement-Based Seismic Design of Structures
413.5 [= 6.0m—ae— 6.0m —4e— 6.0m—+ 3.5m fe
Level 6+ “ ,
Sm Precast Presented Frames
Level 5} Columns 750%300 mma
3.5m Material properties:
Level 4 + f= 40 MPa
atm fuy = 1750 MPa
£ = 400MPa
Level 3-f-] ;
3am Design Drift Limit = 0.025
Level 2 PGA = 0.5g.
3.
Floor Weights: 400 tonnes
Level 1 at each level
Level 0
Fig.5.37 Structure for Example 5.3 (1m = 3.28ft; IMPa =145psi; 1¢ = 2.205kips)
Displacement 4 (mn)
. 55%
0.437 518.7%
i» Period (sec)
40
2.77
Fig.5.38 Design Displacement Spectra for Example 5.3 (Im = 3.28ft,
Solution: The initial data for calculating the substitute structure properties are listed in
Table 5.5. Note that the storey height at level 1 is taken to the mid-height of the beam
and hence is 3.5m, not 4.0m. The design displacements at the various levels are found
from Eq.(5.1) asChapter 5. Frame Buildings 295
(OoS5 oa
where the inelastic mode shape 6 is found from Eq.(5.2) with Hy = 21.0m (68.9f).
Table 5.5 Substitute Structure data for Example 5.3 (m=3.28ft, 1t=2.205kips)
Level,i | Height | Mass m; & A miAy mA? mAHi
Hi (m) | (tonnes) (m)
@ @ 6) @~ | © ©) @. 8)
6 21.0} 320.00 1.000, O.411 1315 54.05 | 2762
5 17.5 400.00 0.880. 0.362, 144.6 52.28 2531
4 14.0 400.00 0741 0.304 121.8 7 1705
tS 10.5 400.00 0,583 0.240, 95.9 22.99 1007
2 7.0 400.00 0.407 0.167, 67.0 11.21 469
+ 3.5 400.00 0.213 0.088 35.0, 3.06 123
Sum 595.8 180.67 8596
Design Displacement: From Eq.(5.3), and Table 5.5, Cols (6) and (7):
Ay = >i (m.&2)) > (mA, = 180.7/598.8 = 0.303m (11.9in)
7 a
Effective Height: From Eq,(5.5) and Table 5.5, Cols (6) and (8):
H,= Ylma,H, VS (m,A,)=8595.6/595.8 = 14.4 (69% of building height)
Yield Displacement: From Eq.(5.71), the yield drift for a hybrid frame is estimated as
0.0005 x 6.00/1.00 = 0.003
6, = 0.00052
A,
We assume a linear yield displacement profile with height, and hence the yield
displacement at the effective height is
A, = 0.003% 14.4 = 0.0432m — (1.70in)
System Duetility: From £4.(5.6) the system ductility is thus
0.303
A, 0.0432Chapter 5. Frame Buildings 299
=1133V,, /6323
These seismic beam shears are listed in Table 5.6 as Col(I4). Note thar these will
subsequently be amplified for flexural overstrength, and to include gravity shears. Note
hat the column storey shears of Table 5.6, Col.(10) should also be increased by 7.8%, to
account for P-A effects, but since it is the ratio of these to the sum of the storey shears,
which is of course unaffected by the increase, the corrected column storey shears are not
listed. It will, however, be necessary to implement the increase when column capacity-
design shears are computed.
Beam Seismic Moments: Since the seismic frames are peripheral, and precast, gravity
ioads will be small in comparison with seismic moments, and slab reinforcement will not
directly add to the flexural strength (though this will depend on the connection details
inerween the slabs and beams. If precast prestressed floor systems with cast i sitr
connections 10 dowel reinforcement in the precast beams are provided, the moment
enhancement can be very significant). tis elected to have equal negative and positive
moment capacity at the column faces. Assuming that the columns are 600X600 mm
-23.6%23.6 in), the beam seismic moments at the column faces are given by Eq.(5.36) as
M sists =V gj (Ly Ne 2 =V 5 (6.0 - 0.6/2 = 2.15,
‘The resulting beam design moments are listed in Table 5.6 as Column (15). As noted
above, 60% of the required moment capacity will be provided by central prestressing, and
40% by mild steel reinforcement. Details of the design process are elementary, and are
nor included here.
Beam Design Shear Forces: Bear flexural design using post-tensioning is likely to
result in lower overstrength factors than with conventionally reinforced sections, since
the range of ultimate rensile strength of the prestressirig steel is typically low. We design
the sections using advanced models for steel and concrete stress-strain curves (see
Section 4.2), and assess possible overstrength for the fraction of moment capacity
provided by the prestressing as 10%, and for the fraction provided by the milé steel
couple as 20%. Assuming a perfect match between required and provided steel areas, the
overstrength factor can thus be assessed as @ = 0.6X1.14+0,4x1.2 = 1.14, We adopt a
value of 1.15,
Gravity loads on beams for the seismic load case, including a 20% vertical dynamic
amplification factor are 8kPa (0.167kips/sq,fi). With a 3m tributary width, the distributed
load on the peripheral beams is thus 24 kN/m. (1.64kips/fi). Modifying Eq,(5.57) by
using the clear span, Z_ between column faces:300 Priestley, Calvi and Kowalsky. Displacement-Based Seis
Design of Structures
joy = 2XLASMy 24x54
“ 54 2
—24x = 0.426M, +64.8-24x
The maximum value of Vy occurs at the column face, and is listed in Table 5.7. Col-(16)
‘Table 5.7 Capacity Design Calculations for Example 5.3
[Tevet [Vee or Moa] Mon | Vous] Vom
| GN) (kNm) | (Nm) | (KN) (kN)
L.@ a6) 7) (is) (19) 20) 1)
fees: 1345 | 1.000 | 209.2 | 4184 | 2588 | 7.6
fae 23 | i400 | 3088 | 6176 | 3312 6624
(es, 2758 | 1912 | 5729 | 1145.9 | 392.2 784.4
3 326.7 | 1.812 | m0 | 14221 | 4402 880.5
2 362.2 1.812 807.5 1614.9 | 4737 947.6
Ed 3808 | 1812 [3579 | 17158 | 4913 982.6
0 1 604.3 | 1388.6 | 491.3 982.6
Column Design Moments: Analysis for column flexure follows the suggestions at the
end of Section 5.5.2(b), where beam moment input to a joint is shared equally berween
the columns above and below the joint, With the overstrength factor of @ = 1.15, as,
calculated above, the overstrength ductility demand, from Eq,(5.62a) is
‘The dynamic amplifications factors for the region from the first floor to the % point
of height, and at roof level, are found from substitution into Eqs.(5.61a) and (5.61b)
respectively as
@,, =1.15+0.13{u? 1)=1.15+0.13%5,09=1,812 and
@,,=10
The full list of dynamic amplification factors for column moment, following the
distribution of Fig.5.25, are provided in Table 5.7 as Col.(17). Column design moments
at the joint centreline are thus found from Eq,(5.60)
0,My 20°@,M, =1.150,M,Chapter 5. Frame Buildings 301
ere Mg are the column moments corresponding to the design lateral forces. In
-ccordance with the recommendations of Section 5.5.2(¢) the beam moments at the joint
centroid are equally distributed to the columns above and below the joint. Note again that
*he beam moments listed in Table 5.6 Col.(15) apply at the columa faces, and have to be
cacreased by 11.1% to obtain the joint centre beam moments. ‘The above equation for
required dependable column moment capacity at the joint centre can thus be rewritten as
O,My 29°, My =1.150/M, =1.150/(.11 Micon !2)
vhere the input beam moment is che sum of moments from beams on opposite sides of
che joint, and the denominator, 2, represents the equal subdivision of moment input to
che columns above and below the joint. As with Example 5.2, the exception occurs at the
zoof, where all the moment is taken by the top of the upper level column, Resulting
Gesign moments at the joint centroid for the exterior and interior columns are listed in
Table 5.7 as Col.(18) and (19) respectively. Note that, as is commonly the case with high
ductility levels, the column reinforcement at level 1 may need to be significantly higher
shan at the columa base, though it must be remembered that the design moments occur
athe level of the top or bottom of the beam, and will be thus be lower than the values
Listed for levels 1 t0 6 by an amount equal to 0.5 Why, where V° is the overstrength shear,
calculated in accordance with the next section..
Column Design Shear Forces: Column design shear forces are determined in
accordance with Section 5.8.5(¢). Equation (5.65) applies. We divide the total storey
column shear between the interior and exterior columns in the proportion 1:2, and hence
the exterior column shears are given by
Mi +ME
C
The resulting shear forces for exterior and interior columns, based only om
amplification of the design shears, are listed in Table 5.7 as Cols.(20) and (21). ‘These
values would need to be checked against the shear corresponding to development of
coverstrength moment capacity at top and bottom of the column, as indicated at the right
of the above equation.
OM oar 2 OV + OM 5 pope) 6 = LAM, + 01K Hp pou! OS
3.12 MASONRY INFILLED FRAMES
3.12.1 Structural Options
Masonzy infill in frames is common in Europe, Asia, and Central and South America.
Although it is less common in North American, New Zealand and Japanese design, it is
still sometimes used in these counties to provide fire resistance between buildings on
boundary frames. However, if the interaction between the frame and the infill is not302 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
properly considered in design and construction, the consequences can be disastrous when
subjected to high intensity seismic action. Examples of struccural failures of infilled
frames in earthquakes are common.
“There are nwo different approaches to construction with masonsy infill. One approach
is to place flexible material between the frame and the infill so that che ewo elements
(frame and infill) can deform independently under the desiga seismic intensity, without
contact, except at the base of the infill, This requires special reinforcement in the infill,
and some means of providing out-of-plane support", ‘The alternative approach is to
build the infill hard up againsc che frame, and to account for the structural interaction in
the design, Generally when infill is built hard up against the frame, the infill is
unreinforced, and susceptible to damage and failure at comparatively low drift angles, The
philosophy adopted in countries where such construction is permitted is that failure of
the infill is non-structural, and can be repaired after the earthquake.
5.12.2 Structural Action of Infill
—
(a) Full-height infill {b) Partial height infill
Fig.5.39 Masonry Infill in Structural Frames
The description of structural action in this section is based on the assumption of
typically weak infill, using hollow tile bricks, as is common in European and South
American countries. In the initial stages of seismic response, infill which is constructed in
contact with the frame on all boundaries, as suggested in Fig.5.3%a), acts as a diagonal
brace to resist frame action. High compression stresses develop in the infill at the ends of
the brace, indicated in Fig.5.39(a) by the diagonal arrows, and separation between infill
and frame occurs at the other ovo corners. Because of the high in-plane stiffness of the
infill, forces induced in che frame members are modified to be primarily axial rather thanprer §. Frame Buildings 303
ral, Stiffness of the frame members (assuming reinforced concrete) is close to the
_scracked value, though the axial stiffness of the column placed in tension by the truss
on of response (the left column in Fig.5.39(a)) may be significantly reduced by
ing, Overall stiffness of the building is very much higher than that corresponding to
ure frame action, and initial periods of response are low.
At comparatively low drift angles — typically 0.003 to 0.005 - initial failure of the infill
urs, generally by crushing of the infill at the contact corners and severe diagonal
ccking parallel to the strut. Sliding shear failures may also initiate along horizontal bed
ors, Subsequent to this, the structural action modifies from a truss or braced-frame
2on to that of a pure frame action, with the infill playing no further structural part in
+e response. This transition is normally complete by a drift of about 0.01
Where the infill is part height, as indicated in the ground floor of Fig.5.39(), the
ser end of the infill brace bears against the column, In the region of column above the
=i shown shaded in Fig, 5.39(b), the structural shear will be resisted by column flexure
era short length, with a potential for soft-storey shear failure developing, particularly if
+s infill has significant strength.
Perhaps the most dangerous condition occurs when the infill is not distributed evenly
sand the structural plan. It has been estimated that in the 1985 Mexico city earthquake,
store than 40% of the building failures occurred to buildings on corner sites, where fire
=-sistant infill was placed only on the two boundaries next to the adjacent buildings,
ating severe torsional ecceniricity!?"l, This situation must be avoided at all costs.
3.12.3. DDBD of Infilled Frames
A pwo-stage design process will be necessary when structural infill is used in
- onstruction. The design philosophy will require that under a serviceability level
acthquake, infill failure does not occur, while under the damage-control earthquake,
same response is acceptable. Normally this will mean restricting drifis under the
iceability earthquake to 0,005, and under the full design earthquake to 0.02 or 0.025,
Zependent on the code drift limit.
The logical procedure will be to design for the damage-control limit stare using
DDBD, on the assumption that the infill has no structural significance. The frame is then
Avemge
— Dasign
L
10009 20000 30004000
‘Moment (kNm)
Fig.5.41 Mean Peak Storey Displacements Response/Overturning Moments
vs. Design Values for Design Example 5.46
STRUCTURAL WALL BUILDINGS
6.1 INTRODUCTION: SOME CHARACTERISTICS OF WALL BUILDINGS
Buildings where the primary or only lateral-force resisting mechanism consists of walls
are frequently called “shear wall” buildings. As has been pointed out elsewherel?", this
has unfortunate connotations, as it implies that response is shear-dominated, whereas the
desired response is ductile flexural action, with shear controlled by capacity design
measures. Consequently we follow the lead of [Pi] and use the terminology of this
chapter's ttl.
The performance of structural wall buildings in recent earthquakes has generally been
200d, and complete collapse under even extreme seismic excitation is rare. Exceptions
sive occutted primarily as a result of foundation inadequacies. A detailed and complete
discussion of the advantages and seismic performance of structural wall buildings is
available in [P1], and only a brief summary will be provided herein
6.1.1 Section Shapes
@) (© @ ©
Fig.6.1 Common Section Shapes for Structural Walls
The choice of possible section shapes for structural walls is limitless, though simple
and symmetrical shapes are to be preferred. Some of the more common shapes are
llustrated in Fig. 6.1. For the rectangular section of Fig.6.1(a), flexural reinforcement may
313314 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
be uniformly distributed along the length, or concentrated in end regions, with only
nominal reinforcement distributed in the central region. Uniformly distributed
reinforcement has the advantage of imparting improved shear resistance, particularly
against sliding shear on the wall base, but results in a lower first-yicld moment than will
occur when much of the flexural reiaforcemnent is concentrated at the ends. The flexural
strength, however, will be lite affected by the distribution of reinforcement, provided
the total area is constant, A similar conclusion was noted for distribution of beam
reinforcement, in Section 5.6.1
“The section of Fig.6.1(b) has boundary elements of increased width at each end of a
rectangular wall section. This shape is often used when beams frame inco the ends of the
wall section, as suggested by the dotted lines. When the wall extends over the full length
of one end of a building, there may also be intermediate boundary elements to
accommodate beams of internal frames extending perpendicular to the wall, on one side.
Ir should be noted that the structural system implied by this, of end walls providing
seismic resistance in one direction, and frames in the perpendicular, and longer direction
can result in undesirable seismic response. Under diagonal attack, the boundary element,
which is essentially a column, at one end of the wall may be subjected to compression
stresses close to the concrete compression strength from the cantilever action of the wall,
while being deformed laterally by frame action in the orthogonal direction. The high
compression siress in the boundary element reduces its moment capacity in the frame
direction, and flexural yielding of the boundary element may result, Local P-A effects can
become critical. The combined wall and frame action on this boundary element at levels 1
and 2 can result in instability and collapse of the end region of the wall, as was observed
with several apartment buildings after the 1995 Kobe earthquake.
Fig.6.1(6) shows a T-section wall, which is common in buildings with internal central
corridors, such as hotels and apartment buildings. In these cases the flanges form part of
the corridor wall between doorways, and the web divides different hotel rooms or
apartments. ‘The behaviour of T-section walls in the direction parallel to the web is
characterized by different strength and stiffness in the ovo possible loading directions,
with the wall generally being stiffer and stronger when the flange is in tension than when
it is in compression, The yield curvatures in the opposite directions, may also differ (see
Section 4.4.3(c)). Frequently buildings will contain identical, but anti-symmetzical T-
section walls when rooms are symmetrically placed on either side of a central corridor.
‘This simplifies the structural characterization, since the average values for strength and
stiffness for flange in tension and flange in compression can be adopted, and total system
strength and stiffness will be the same in both directions.
Finally the C-section wall of Fig.6.1(@) is common when walls enclose a service core
of lifts, stairs and possibly toilets, and is often combined with a symmettically opposed C-
section, as suggested by the section shown in dotted outline, These sections will often be
connected by coupling beams, shown in Fig.6.1(¢) as dashed lines, resulting in coupled-
wall behaviour, discussed later in this chapter, in Section 6.8. As with the T-section walls
of Fig.6.1(¢) strengths and stiffness of a C-section wall differ depending on the directionChapter 6. Structural Wall Buildings 315,
of response, and when loaded parallel to the web, torsional response must be expected
unless a balancing symmetrically opposed element is present, as suggested in Fig.6.1(c)
6.1.2 Wall Elevations
(a) Cantilever Wall (b) Wall with Openings —_(c) Coupled Wall
Fig.6.2 Categories of Structural Walls
The three main categories of structural walls are illustrated in Fig.6.2. Only two are
suitable for seismic resistance. The cantilever wall of Fig.6.2(a) is the simplest, and the
most straightforward in terms of predicting seismic performance. Provided proper
tention is paid to dynamic amplification of moment and shear, inelastic action occurs in
\ flexural plastic hinge forming above the base of the wall, and extending some distance
sp the wall, as indicated by the shaded area, Above this region, the wall remains elastic.
The second category, shown in Fig.6.2(b) is a wall with openings, where the openings
we insufficient to provide frame-like action. In the example shown, the piers beeween
penings are smaller than the beams above and below the openings. With the
>toportions shown it is very difficult to avoid inelastic action occurring by flexural
clding or shear failure in the piers, generally below the first floor, as indicated by the
shaded areas. This form of construction is unsuitable for seismic resistance unle
response can be assured to be elastic, or near-elastic (displacement ductility demand less
shan 1.3)
Coupled walls, shown in Fig.6.2(6), are designed to form flexural plastic hinges at the
all bases and in the coupling beams. These provide an efficient mechanism for resisting
scismic forces, with reduced displacements. These are separately considered in Section
8
6.1.3 Foundations for Structural Walls
Foundations for structural walls are generally rather massive to enable the overturning
moment ta be resisted by gravity effects. This is illustrated in Fig.6.3, where the316 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
overstrength shear V® acts at an effective height of He above the wall base, The
overturning moment at the base of the wall is approximated by
y
ME =V°H, =(A, 0 +Wy)-l, |
(1)
where Wy is the weight, including self weight, supported by the wall at the base. Ay is the
total area of flexural reinforcement in the wall, with overserength yield stress of fy’. The
criterion for stability of the wall on its foundation under ductile overstrength response is
given by
Mg=(Wy +W)(Lp~a)/2> Me =V°(H, + hp) (2)
where any passive resistance of soil on the face of the footing is neglected, and Ly is the
footing length, Wr is the footing weight, and a is the length of the footing in
compression at ultimate soil bearing strength, Pay given by
W, +W,
a= w tte 63)
Pow Be
and where Bp is the footing width.
Fig. 6.3 Overturning Resistance of a Cantilever Wall
Equation (6.2) applies for a spread footing. ‘The size of the footing can be reduced by
supporting the footing on piles with tension capacity, The modification to F.q.(6.2) for
the case where some overturning resistance is provided by tension piles follows the sameChapter 6, Structural Wall Buildings 317
principles as outlined above, and is not developed further herein. Overturning tesistance,
Mpg for walls is discussed further in Section 6.5.
6.1.4 Inertia Force Transfer into Walls
Since there will generally be much fewer walls in a structural wall building than
columns in a frame building, floor inertia forces to be transferred into the walls can be
large. To ensure satisfactory force transfer, special consideration must be given to internal
force transfer within the floor diaphragm, and particularly to force transfer between the
floor diaphragm and the wall. This is normally effected by shear-friction action, It should
be noted that floor diaphragm inertia forces are strongly influenced by higher mode
action, and may be many times higher than the level predicted by the design
distribution of storey force corresponding to the inelastic first-mode response, as used in
the DDBD procedure. Higher-mode effects are discussed in detail in Section 6.6.
6.2. REVIEW OF BASIC DDBD PROCESS FOR CANTILEVER WALL.
BUILDINGS
‘The fundamentals of DDBD were introduced in Chapter 3, with reference to Fig.3.1
and were summarized in Section 5.2.1 for frames. Most of the equations are identical for
wall buildings, and hence are not repeated here. Equivalent SDOF design displacement is
given by Eq.(5.3), for equivalent mass by Eq.(5.4), for effective height by Eq(5.5), for
design ductility by Eq.(5.6), for effective stiffness of the SDOF structure by Eq.(5.19)
and for design base shear by F.q.(5.11). The base shear is distributed to the floor levels in
accordance with Eq.(5.12). Differences primarily relate to the deflection profiles and
damping values to be adopted for design. ‘These are discussed further here related to
cantilever walls. Coupled walls are separately considered in Section 6.8
6.2.1. Design Storey Displacements
(a) Yield Displacement: in Fig.3.18, the curvature profile up the wall was represented
by a straight line from the yield value at the base to zero at the top of the wall. This may
appear incompatible with the moment distribution corresponding to distributed lateral
seismic forces, which results in a curved moment pattern, and since curvature is
essentially proportional to moment, the curvature distribution should similarly be curved.
Further, in the upper regions of the wall, the moments may be less than the cracking
moment, and gross-section curvatures will be much less than cracked-section curvatures.
This is illustrated in Fig.6.4, where the “design forces” curvature profile at yield
corresponds to an inverted triangle distribution of lateral forces, and may be compared
with the suggested linear distribution. In both cases, curvatures have been put in
dimensionless form, dividing by the yield curvature at the wall base. Aiso shown in
Fig.6.4 is the curvature distribution where the wall is considered to be uncracked for theChapter 6, Structural Wall Buildings 324
limit strain was based on recommendations of Section 4.2.5 using the following
information: Wall width 250mm (9.8in), transverse reinforcement 10mm (0.393 in) at
100mm (3.94in) centres across the wall in the end region, with 10mm bars along each
side. Transverse reinforcement at $= 100mm (3.94in) vertically (see Fig.6.6)
Fig.6.6 Wall-
ty and Damage-Control Analyses
Area ratios of confinement in both directions are equal at
4s. 785 _ 9.9975
Par = Po =H 100x100
Adopting a confinement effectiveness coefficient of C, = 0.5 for walls, as recommended
in Section 4.2.2, the effective volumetric confinement ratio is thus, from E.q.(4.7)
Po = Pa + Pa, = 2(0.5x0,00785)= 0.00785
From Eq.(4.21), taking the confined compression strength as 1.4f%, and the strain
capacity of the transverse reinforcement as 0.12, the ultimate compression strain is
6, =0.004-41.4PLném. — 9,904 41.4 0200785%450%0.12 _ 9 o1g
- 1.4x30
The strain limit for the longitudinal reinforcement is based on 0.6&, as recommended in
Section 4.2.4(c), where the steel steain at maximum stress is 0.10. The results in Fig.6.7
were calculated based on expected concrete cylinder compression strength of fie = 30
MPa (4.35ksi), and reinforcement yield strength of fe = 450 MPa (65ksi), with an
ultimate/yield strength ratio of ff, = 1.3. However, as the data are expressed in
dimensionless form, the results will be applicable to other concrete and reinforcement
strengths within reasonable variation (say 20MPa 4 6 Rt 6 6s 68 72 76 8
Aspeee Ratio (A, =H, fy) Moment Magnitude
(0) SDOF Vita Displacement (6) Plateau Displacements
Fig.6.9 Wall Yield Displacements for €, = 0.00225 Compared with Plateau
Displacements
Fig.6.9(a) plots system yield displacement for walls of different aspect ratio and height,
based on Eg.(6.15) for a yield stress of 450MPa (65.3 ksi). These can be compared with
the plateau displacements presented in Fig.6.9(). An example of the use of this figure is
shown by the dashed line in both figures. For a 16 storey wall with uniform storey heights
of 3 m, (9.84 ft) and a wall length of 8 m (26.2 fi), the total height is 48 m (157.4 ft) and
the aspect ratio is 6. Figure 6.9(a) is entered at A, =6, and the system yield displacement
of 0.292 m (11.3 in) is transferred across to the seismological data of Fig.6.9(b). From this328 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
it is seen that for a causative earthquake of My = 6.5, the wall will be expected to
respond clastically if the design PGA is less than 0.3g.
6.3.3 Multiple In-Plane Walls
Aspects related to displacements and damping when several walis of different length
are in the same plane or in parallel planes have been discussed in Section 3.5.4, and
illustrated in the design example of Section 3.5.5. Aspects to be considered are the
different ductility demands, and hence different equivalent viscous damping values for
each wall, the distribution of strength between walls, and the global (system) damping. It
was pointed out that the distribution of strength between walls is purely 4 designer’s
choice, uninfluenced by considerations of initial elastic stiffness, and that 4 common,
logical choice will be t0 use equal flexural reinforcement ratios in all walls, leading to the
base shear force being distributed to walls in inverse proportion to the square of the wall
length. See also comments in Section 6.2.1(d).
6.4 TORSIONAL RESPONSE OF CANTILEVER WALL BUILDINGS
6.4.1 Elastic Torsional Response
A brief summary of consideration of torsional effects in DDBD was presented in
Section 3.8. However, since the topic has special relevance to structural wall buildings, as
a result of the potential for torsional eccentricity with building plans containing walls of
different lengths, the topic is treated in greater detail in this chapter.
We start with consideration of elastic torsional response, which though not directly
relevant to inelastic structural response, will be referenced and modified to develop an
appropriate design approach. Until recently, torsion has been treated as an elastic pheno-
menon. In the elastic approach shears resulting from distribution of the clastic base shear
force between different lateral-force resisting elements are modified to include shear
forces induced by twise of the building resulting from eccentricity between the centre of
mass and the centre of elastic stiffness of the building. Typically this eccentricity is
augmented to include “accidental” eccentricity resulting from uncertainties in distribution
of mass, and in material properties, which hence result in uncertainty in the positions of
the centre of mass and the centre of stiffness,
The approach is illustrated by refereace to the plan representation of the wall building
shown ia Fig. 6.10, which has plan dimensions Ly and Lz (¥ is the building vertical axis)
Walls 1 and 2, of different lengths, resist seismic actions parallel to the Z axis, and walls 3
and 4, also of different lengths, resist seismic actions parallel co the X axis. The centre of
mass, Cy is assumed to be at the centre of the floor plan. ‘The origin of the axes, raken at
Caw is thus also in this case at the centre of the floor plan,
The centre of rigidity, or stittiness Cr is displaced from the centre of mass by distances
eax and egz in the X and Z directions respectively. Note that since stiffness of structural
elements depends on strength, as established in Section 4.4, the actual stiffness330 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Vizg =Vese,2€ax (6.20)
Téa base shear Vax acts concurrently in the X direction, then additional shear is
induced in wall # by the additional torque Vigase,x-€rz:
k(x, -e,
Vox =Veoseax rz balou) (6.206)
a
‘The total shear carried by the elastic wall is thus
Vi Van +Vizg *Vixa (6.208)
There are a number of ctiticisms chat can be levelled at this classical approach to
torsional response of structures. First, as with all aspects of force-based design, it relies
On initial estimates of element stiffness, to determine both the centre of stiffness, and the
torsional stiffness. As has been repeatedly pointed out, element stiffness is proportional
to strength, and hence significant errors can be expected unless an iterative design
approach is used to modify element stiffness after initial estimates of strength have been
obuined. Typically, this is not done, and also, typically, member suiffnesses are
significantly in error in conventional force-based design.
More importantly, the stiffness approach does not account for changes in
performance once one or more of the walls yield, after which stiffness considerations
become largely irrelevant, nor does it take into account the very\significant role of
torsional mass inertia in modifying the torsional rotation implied by Eqs.(6.20b) and
(6.20c). As a consequence, the elastic representation of torsional effects cannot be used
to predict the torsional response of ductile systems. This has been recognised for some
time, and early considerations of ductile response of systems including torsional
effects(.2P1 distinguished between so-called torsionally restrained and torsionally
unrestrained systems. The distinction between these categories is clarified in Fig.6.11.
For seismic action parallel to the Z axis, the two transverse walls in Fig.6.11(a), which
are in line with the centre of mass Cw, can play no part in resisting any torsional moment.
Thus if one or both of the walls (1 and 2) parallel to the Z axis yield, the torsional
stiffness will drop to a very low value, dependent primarily on the post-yield stiffness of
the yielding walls, Consequently, in carly studies based only on static torsional
mechanisms, it was expected that if one wall yielded before the other, the centre of
stiffness would shift to coincide with the remaining elastic wall, and essentially
unrestrained rotation would result. The system is thus classified as torsionally
unrestrained, and earlier texts (e.g, [P1]) advised against such systems.Chapter 6. Structural Wall Buildings 331
1 °
Cu
1S 0 SSS 2
Cu
LZ
Fasc
(a) Torsionally unrestrained (b) Torsionally restrained
Fig.6.11 Different Wall Layout for Building Plan
“The structure of Fig.6.11(b) has boundary elements on all four sides, and thus if walls
1 and/or 2 yield under seismic force parallel co the Z axis, the structure retains torsional
stiffness, albeit reduced from the initial elastic value, as a result of the lever arm between
the two transverse walls. The system is thus termed torsionally restrained, and was initially
felt to be structurally more desirable than unrestrained systems.
Recent research by Castillo and Paulay'tl, followed and extended by Beyerl*!, have
shown that torsional inertia plays an important role in modifying structural response of
both unrestrained and restrained systems, but is difficult to quantify accurately, since peak
translational and torsional response do not occur simultaneously. Both studies carried out
extensive inelastic time-history analyses to investigate ductile response including torsional
effects. The initial studies by Castillo and Paulay investigated 2D plan simulations of wall
buildings. Beyer extended these studies, and also carried out a number of full 3D analyses.
The following two sections present a brief summary of their findings.
6.4.2. Torsionally Unrestrained Systems
The studies of torsionally unrestrained (TU) systems by Castillo and Paulayl¢i
showed that the modification to response resulting from torsional mass inertia was
considerable. Considerations of static equilibrium would indicate that if we assume that
the centre of mass Cy is at the centre of the building plan, and further assume that wall 2
in the TU system of Fig.6.12 is weaker, as well as less stiff than wall 1, then it will yield
prior to wall 1, since the ratio of forces developed in the two walls is defined by geometry
Thus, again assuming that Cys is at the centre of the building plan, the forces in the two
walls must be equal. For example, assume that the nominal strengths of the walls are
related by Vyy=1.4Vy2; then when wall 2 yields, wall 1 will have an equal force,
corresponding to 0.714Vyy. Since the force in wall 2 will increase only slightly (due to
post-yield stiffness of the force-displacement response) as the structure responds in the
inelastic range, the force in wall 1 would never be expected to reach yield, and all system
ductility would result from inelastic response of the weaker wall 2. This would imply an332 Priestley, Calvi and Kowalsky.Displacement-Based Seismic Design of Structures
expected ductility demand on wall 2 close to twice the system ductility demand, measured
at the centre of mass. The corresponding displacement profile is shown in Fig.6.12 by
the dashed line A,
Vs
Maximum, ignoring,
}— torsional inertia
Maximum, including
torsional ine!
Displacement
Fig.6.12. Displacement Response of a TU System
‘The inelastic time-history analyses of Castillo et al showed that this behaviour did
indeed occur, but only if the torsional mass inertia was set to zero. When realistic values
for torsional inertia were included in the analyses, it was found that the stronger wall
would always yield, provided system ductility demand was significant, and would be
subjected to considerable ductility demands, In fact the stronger wall would normally
vield before the weaker wall, provided that the stronger wall was also stiffer than the
weaker wall, as will usually be the case. The corresponding displacement profile is shown
in Fig.6.12 by the solid line B. However, this line is a little misleading, as the peak
displacements for walls 1 and 2 do not necessarily occur at the same time, nor does the
peak rotational response occur simultaneously with cither peak displacement.
Castillo et al recommended that minimum required strengths for elements of TU
and also of TR) systems should be based on considerations of static equilibrium. Thus,
with reference to Fig. 6.12 the minimum nominal strength of walls 1 and 2 would be
related to the total design base shear Vpave by:
Vax 2Viggce (6.21)Chapter 6. Structural Wall Buildings 333
Excess strength, above that required by Eq.(6.21), for either wall would result in
strength eccentricity ey, (see Figs. 6.10 and 6.12) where
V,
ok,
V+¥,
ey x (6.22)
Normally excess strength would result from the longer wall needing less than the
specified minimum reinforcement ratio to provide the required flexural strength
corresponding to V;, and hence actual strength would exceed required strength. Note
that in accordance with capacity design principles the flexural strength will be matched to
the moment corresponding to the required shear strength (see section 4.6). Castillo et
all found that excess flexural strength in cither wall did not adversely affect the
performance of TU systems. For the normal case where the longer wall (wall 1) had
excess strength, the displacement demand on both the centre of mass and the longer wall
was, as expected, found to be reduced by the strength eccentricity resulting from excess
strength. The displacement demand on the shorrer, more flexible wall (wall 2) was found
to be almost independent of the excess strength ‘and hence of the strength eccentricity)
when the excess strength occurred in wall 1 only. Excess strength in wall 1 could thus be
used to reduce the displacement demand on wall 1, without adversely affecting wall 2.
Note however, that strength eccentricity is expected to increase torsional displacements
as will be demonstrated shortly.
Tt will be noted that it may not be immediately clear which of the walls will govern the
design. Normally, with different wall lengths as suggested in Fig.6.12, the ductility
demand on the longer wall will be higher than on the shorter wall, which is contrary to
expectations based on static equilibrium. Hence if material strains govern design, the
longer wall will be critical. On the other hand, the shorter wall will be subjected to larger
displacement demands, and hence if drift governs design, the shorter wall will be the
critical element.
The conclusions of Castillo et all* for TU walls can be summarized as follows:
‘* Nominal strength of any element should be not less than that required by static
equilibrium considering zero strength eccentricity, for the chosen base shear
force, regardless of the stiffness eccentricity
© Strength eccentricity does not adversely affect the performance of TU systems,
provided it only results from excess strength of one or more elements.
* If these guidelines are met, the displacements will not be greater than those
estimated for zero strength eccentricity.
© The results are rather insensitive to the magnitude of torsional inertia, provided it
is within reasonable bounds. (Castillo et al investigated variations of $20% from
a uniform mass distribution. Higher values, which are more probable than lower
values, reduced twist).334 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
6.4.3, Torsionally Restrained Systems
B
N
SS
| ‘Wall 3
+ ry = 25 m.
Fig.6.13. Torsionally Restrained Building with Strength and Stiffness Eccentricity
for Z direction response (after Beyer!*4))
Castillo et all€*! and Beyer"! both studied torsionally restrained (FR) systems, and,
with a few exceptions, obtained generally similar results. The following comments are
based on the more extensive studies by Beyer. These included 2D plan simulations of
cight-storey buildings, and 3D studies of 2, 4, and 8 storey buildings with strength and
stiffness eccentricities. Analyses were carried out with a suite of five avtificial
accelerograms spectrum-matched to a displacement spectrum linear with period to a
period of 5 seconds. It was found that the spectral reduction values for levels of damping
higher than 5% agreed closely with the “old” EC8 equation given by Eq,(2.8). Beyer’s 2D
analyses were based on a floor plan of 25m X 15m as shown in Fig,6.13, with the lengths
of walls 1 and 2 being 8m and 4m respectively. The two transverse walls (3 and 4) cach
had lengths of 5m, and similar total system strengths were provided in both X and Z
ctions. A base-level design was chosen to provide an average drift of 0.02, with zero
strength eccentricity, using displacement-based design procedures, Thus the strength
provided to each wall in the Z direction was the same for the base-level design, despite
the different wall lengths. Since the yield displacement for the walls would be in inverse
proportion co wall length (see Section 4.4.3), this resulted in wall 1 having twice the elastic
stiffness of wall 2. Thus, though the strength eccentricity for the base-level structure was
ev = 0, the stiffness eccentricity was €g = 0.167Lx. A series of analyses was carried out
increasing the strength of the long wall relative to that of the shore wall by ratios of
1.0SAS1.8 where A=Vi/V>, increasing both the stcength and stiffness eccentricity. This
was effected in ewo ways. In the first, the strength of wall 1 was increased while keeping
the strength of wall 2 constant. The total system strength thus increased by a maximumChapter 6, Suructural Wall Buildings 335
of 40%. In the second case, the strength of wall 2 (the short wall) was reduced as wall 1
strength was increased such that the total system strength remained constant. In both
cases, the stiffness and strength eccentricities, with kz/=2Akz», are found from Fgs.(6.18)
and (6.22) as:
_0.5(1-24)
~ 1424
050-4)
(1+)
av y (@) and eyy Ly &) (6.23)
Results of the analyses, averaged over the 5 accelerograms, are summarized in
Fig.6.14. It will be noted that despite the zero strength eccentricity at A=1.0, some
torsional response is evident, as a result of the stiffness eccentricity, with displacements
of the short and long walls being about 10% larger and smaller respectively than the
centre-of-mass displacement. As the strength of the long wall increases, resulting in total
system overstrength (Fig.6.14(@)), displacements of the centre-of-mass decrease, but the
displacement of the short wall remains almost constant.
04
E 0.3 =
z z
goa i
2 2 '
e & '
Zot a :
° oF t T t41
1 1.2 14 16 18 1 12 14 1.6 1.8
‘Wall Strength Ratio (V1/V2) Wall Strength Ratio (V1/V2)
(@) Total Strength Increases (b) Constant Total Strength
Fig.6.14 Displacements of TR Wall Building with Strength and Stiffness
Eccentricity (after Beyer!™I)
‘When the total system strength is kept constant by reducing the strength of the short
wall as the long wall strength increases, the centre-of-mass displacement remains
essentially constant, while short and long wall displacements increase and decrease
respectively, Apart from local variations which can be attributed co inevitable scatter
resulting from the time-history analyses, the displacements appear to vary iis
the wall 1 overstrength factor. Note that at A = 1.8, corresponding to stiffness and
strength eccentricities of 0.28Lx and 0.143.Ly respectively, the displacements of the walls
differ by about +40% from the centre of mass displacement.
ly with336
Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
From the analyses of Beyer‘! and Castillo et alll the following conclusions can be
drawn:
Displacements of the centre-of-mass can be reliably estimated from a SDOF
model based on the DDBD principles claborated in this text. That is, the
torsional response does not affect the centre-of-mass displacement.
As with TU systems, increasing the strength of an element above that required to
satisfy displacement demands in the structure will not result in incteased
displacements in other structural elements, provided the strength of those
elements are not reduced
“The results are largely insensitive to the value of rotational inertia, within typical
expected variations from the value corresponding «© uniform distribution of
mass.
Varying the strength and stiffness of the transverse walls (Walls 3 and 4,
Fig.6.13), multiplying by factors of 0.5 and 2.0 had only minor effect on the
rotational displacements, with displacements increasing only slightly for softer
transverse walls,
Allowing the transverse walls to yield under the rotational displacements also had
only minor influence on the displacements of the long and short walls.
Increasing the seismic intensity while keeping the system strength constant (Le.
increasing the system ductility demand) had little influence on the magnitude of
twist associated with peak displacement demands when the strength eccentricity
was zero. When appreciable strength eccentricity existed, the displacements of
the walls increased almost in proportion to the displacement increase at the
centre-of-mass. That is, the torsional component of peak displacements
increased with ductility
Eccentricity of mass from the geometric centre of a building plan did not
significantly affect response, provided stiffness and strength eccentricities were
measured from the actual centre of mass,
For buildings with eccentricity about only one axis (as, for example, in Fig.6.13),
seismic excitation directed at a skew angle to a principle axis resulted in reduced
peak displacements in the directions of the principal axes, and hence was not
critical.
For buildings with strength eccentricity about both principal axes (e.g: Fig.6.10),
slight increases in displacements in the direction of the principal axes were
possible under skew attack compared with values resulting from excitation in the
direction of the principal axis. This increase was generally small (less than 10%).
When the effective secant-stiffness perind of the structure (at maximum.
displacement response) was larger than the corner period of the displacement
spectrum (see Section 2.2.2) response was complicated by the fact that centre-of-
mass displacement did not decrease if excess strength was provided to the stiffer
wall, Hence displacements of the flexible wall increased. ,Chapter 6, Structural Wall Buildings 337
6.4.4 Predicting Torsional Response
As noted above, DDBD provides an excellent estimate of the displacement response
of the centse of mass of a corsionally eccentric building. It remains, however, to
determine the displacements at the edge of the building, as effected by torsional rotations.
No exact simplified method of analysis seems possible, as the time-history analyses have
clearly shown that both strength and stiffness eccentricities affect response, and that peak
displacements of centre of mass and the walls st opposite ends of the building do not
occur at the same instant of dynamic response. ‘The following approach, however, which
is consistent with the principles of DDBD has been found ro provide displacements of
the building edges in close agreement with results of time-history analyses for both TU
and TR systems.
The maximum displacements of a building plan can be approximated by a translation
of the centre of mass, determined by DDBD principles plus a nominal rotation Ay where
8y = Vase Ce! Feu (6.24)
where Vause is the design base shear force, ex is the elastic stiffness eccentricity, given by
Fiq.(6.18), and the ductile rotational stiffness, Jaw is modified from the elastic rotational
stiffness Jp of Eg,(6.19), dividing the wall stiffness in the direction considered by the
system ductility fly:
+ Yiu, ~enz) (6.25)
7
Note that the elastic stiffness is used for wall elements perpendicular to the direction
considered. Thus the effective stiffness of wall elements in both X and Z directions is
used, since only the Z direction walls are expected to be subjected to significant ductility
demand under Z ditection excitation,
The displacements of the end walls (Fig.6.12 for TU systems or Fig6.13 for TR
systems) are then found from
A, = Bey + Oy (%; ~ ey) (6.26)
Note that in Eq.(6.26) the strength eccentricity, rather than the effective stiffness
eccentticity has becn used. Similarity to the procedure outlined above for elastic torsional
response will be recognized; the difference being that torsional stiffness is based on
effective stiffness of the elements, and torsional displacement increments are based of
distances from the centre of strength, Cy.
‘The approach outlined above has been checked against analytical results from Castillo
et all and Beyer"™l, and found to give generally good agreement for both TU and TR
systems, For TR systems with stiffness eccentricity but no strength eccentricity it was338 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
found to accurately predict the incremental displacements resulting from torsion for
levels of ductility between 1.3 and 7. Comparisons with average time-history results of the
predictions for short wall and long wall displacements for TU and TR systems are shown
in Figs 6.15(a) and (b) respectively. The TU systems of Fig.6.15(a) refer to the data of
Castillo et all, and were designed for an initial ductility of faye =
eccentricity (i
35 at zero strength
V, =V2). The ratio of wall lengths was fy; = 1.4ly2. The strength of wal)
1 was then increased up 10 a maximum of 1,86V> without changing the wall lengths, or
modifying the strength of wall 2. The increased system strength thus resulted in a
reduction of the centre-of-mass displacement, as is apparent in Fig.6.15(a)
The TR systems of Fig, 6.15(b) refer to Beyer’s data™l, and were designed for centre-
of-mass system displacement ductilities of approximately flys = 3.0. Wall 1 was evice as
fong as wall 2, and strength ratios up to Vj = 1.8V3 were considered, while keeping the
total system strength constant. In each case the average results from the inelastic time-
history analyses are plotted with data points and solid lines, and the predictions using the
approach outlined above are shown as dashed lines. In both cases the centre-of-mass
displacements have been used as the datum, with variations from this for the short and
long walls calculated using Eqs. (6.24) to (6,26), Ir will be noted that the agreement is
satisfactory over the full range of data, with the predictions tending to be slightly
conservative (i. high) for both short and long walls.
z
E03 -s-----b---' E
z 'Short Wall 1 ¢
go2 1 €
a : 3
= 1
#ot ay
OT 94
1 12 14 16 1.8, 2 1 1.2 14 1.6 18
Ratio of Wall strengths (V1/V2) Wall Strength Ratio (V1/V2)
(a)TU system, system strength increases (b) TR system, constant strength
(Castillo’s data)Is (Beyer’s data)!®41
Fig.6.15 Comparison between Predicted Displacements (dashed lines) and
Average Time-History Results (solid lines+data points) for TU and TR systems
Note that the TU and TR structures of Figs.6.15(a) and 6.15(b) are unrelated, with
different masses, strengths and design ductilities, and no conclusions can be arrived at by
comparison between the two figures.Chapter 6. Structural Wall Buildings 339
6.4.5 Recommendations for DDBD
Again it is emphasised that the best design solution will be to eliminate strength
eccentricity. As discussed subsequently, in such cases, the displacements due to twist are
likely to be less than 10% of the system translational displacement, and thus in most
practical cases could be ignored as well within the expected uncertainty of response.
However, this approach may not always be feasible, and the following design approach is
intended to provide a systematic approach, completely consistent with the DDBD
philosophy of achieving a specified displacement limit state. A simplified approach, with
additional conservatism is briefly discussed in Section 6.4.7.
(4) Design when T. Tc: It was noted in the summary of Section 6.4.3
that torsional response appears to be more severe when the effective period exceeds the
corner period of the displacement spectrum. In such cases, the procedure outlined above
could be non-conservative, resulting in displacements of the flexible wall that exceed the
limit state values. Research is on-going into this behaviour. Until definitive
recommendations are available, it is conservatively suggested that the corner period be
ignored, and the displacement spectrum be continued linearly up to the effective period.
‘This will mean designing for a higher base shear than would result from designing to the
plateau displacement applying for Te > Te. Since there is uncertainty associated with the
correct value for Te as noted in Chapter 2, such an approach is doubly prudent,
(©) Consideration of Accidental Eccentricity: As noted in Section 3.8.1, design for
accidental eccentricity is likely to be ineffective, since it involves increasing the strength of
all structural elements, which simply results in an increase in the torsional moment.
Although the overall consequence will be a reduction in displacements, the effect will be
minor. Hence we do not recommend consideration of accidental eccentricity in DDBD.
(d) Design for Bi-directional Eccentricity: Thus far, the discussion of torsional
response has assumed that the direction of seismic attack is parallel with one of the ovo
principal axes of the structure, In fact, there will normally be excitation components in
both principal directions simultaneously, with the resuleant inertia force acting at an angle
to the structure principal axes, as suggested in Fig.6,10 and 6.17. The example of Fig.6.17
shows a plan view of a building braced with four boundary walls of different sizes,
resulting in strength eccentricity in both principal directions. The inertia force Vy acts
through the centre of mass, Cx, with the centre of strength, Cy, eccentric by a distance ey
measured perpendicular to the line of action of the inertia force.
It is clear that there is a torsional moment acting on the building, given by Viev. It is
also clear that if the seismic intensity is sufficiently high, all four walls will develop their
flexural strength. In this case the inertia and resisting forces are given by:
y, M4, +40) (6.39)Chapter 6, Structural Wall Buildings 345
Wall 3
Wall 2
Wall 4
Fig.6.17 Diagonal Seismic Attack, Plan View
where Vi, V2, Vs, and Vy are the base shears corresponding to flexural strength of walls 1
to 4 respectively. Note that if all walls yield, the directions of the inertia and resisting
forces ate fixed, and are not necessarily parallel co the resultant ground excitation
direction
Given that all four walls have yielded, there is no further torsional resistance available
to limit rotation under the torsional moment. However, torsional mass inertia, resulting
from the distributed nature of the mass across the floor plan will limit the rotation, as
with the torsionally unrestrained case discussed in Section 6.4.2. It would seem that an
estimate of maximum feasible oration should be available by determining the dis-
placement response in the diagonal disection, and hence the ductility in the ewo
orthogonal directions. The effective rotational stiffness can then be determined from a
modified form of Eq.(6.25):
(6.40)
where fy and fiz are the average (system) displacement ductility demands in the X and Z.
directions respectively. The procedure fordetermining expected displacement response
for a designed structure by DDBD principles is covered in Chapter 13.
In fact, it will rarely be necessary to carry out these calculations for new buildings.
Since the strength in the diagonal direction will be greater than in a principal direction
(rypically by about 40%), displacements in the diagonal ditection of the centre of mass
will be less than the displacements in the principal directions under orthogonal excitation
parallel to the principal direction considered. ‘The component of the diagonal
displacement in the principal directions will be even smaller, resulting in a large reserve in
displacement capacity to allow for torsional rotation. Consequently, torsional rotation
under diagonal excitation will not normally be considered in the design process.346 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
6.4.6 Design Example 6.1: Torsionally Eccentric Building
=
+ ai
belo in &
20m) Bm) im) BLO883 fe,
a ‘
rr :
= ng7 so
>| bu 6m) Period (see)
Ly 5m) ed
(@) Floor Plan (b) Design Displacement Spectra
Fig. 6.18 Data for Example 6.1
‘The building shown in plan in Fig.6.18(a) has six storeys, with a uniform storey height
of 2.8 m (9.2 fi), and equal floor weights of 3000kN at cach level, including the roof. The
structural system consists of boundary walls, as indicated in Fig.6.18(a), with internal
prop-columns and flat-slab floors which do not contribute significantly to the lateral
resistance in either of the principal directions. The building dimensions are Ly = 25 m
(82 ft) and Lz = 20 m (65.6 ft). The ewo walls in the Z direction have lengths fy = 8 m
(26.2 ft) and typ = 4 m (13.1 fi). Wall widths of 250mm (9.8 in) are selected. The
difference in wall length results from wall 1 being on a boundary adjacent t other
buildings, while wall 2 is on a road frontage, where minimum disruption to access is
desired. In the X direction the structure is symmetrical, wich two walls of 6 m (19.7 f)
length. The building is to be designed to a damage-control limit state, for which the code
drift limit is & = 0.025. Specified material strengths are f°: = 30 MPa (4.35 ksi) and f, =
420 MPa (60.9 ksi). e flexural reinforcing steel will be 20mm (0.79 in) diameter
tempcore steel with a ratio of ultimate to yield strength of fi/f, = 1.25, and a strain at
ultimate strength of 0.10.
The structure is to be constructed in a region of high seismicity, corresponding w a
PGA of 0.6g, with the displacement-spectrum for 5% damping given in Fig.6.18(b).
Displacement reduction for damping conforms to Eq.(2.8).
Solution: The design process follows the procedure outlined in Section 6.4.5(a)Chapter 6, Structural Wall Buildings 383
a minimum base shear force may not justify the additional design effort. In such cases a
conservative design approach may be adopted.
When the flexible wall governs the design (see Section 6.4.5(a) Step 2), the process is
straightforward, and no simplification seems warranted. However, when the stiffer wall
governs the design, then a reasonable simplification will be to design for 2 centze-of-mass
displacement equal to 1.1Ascim, determined at the effective height, based on the stiff wall
displacement profile. The design steps are then identical to the case when the more
flexible wall governs design.
6.5 FOUNDATION FLEXIBILITY EFFECTS ON CANTILEVER WALLS
6.5.1 Influence on Damping
“The influence of foundation flexibility effects ia DDBD was briefly introduced with
specific reference to cantilever walls in Section 3.5.4(b), where ic was noted that
foundation flexibility increases the elastic displacements, but has a lesser, or zero
influence on the design displacement, depending on whether the design displacement is
strain-limited or drift-limited. In both cases the design system ductility demand is reduced
by foundation flexibility, and as a-consequence, the effective damping may also be
reduced, ‘The topic was also introduced in Section 1.3.4(c) with reference to problems
considering foudation flexibility within a force-based design environment.
In Fig.6.19, the elastic displacement at the effective height resulting from foundation
flexibility is Ap, increasing the yield displacement from Ay to AY. If the design
displacement Ap is strain-limited, then the desiga displacement also increases by
essentially the same amount to A’p. A small additidnal incréase may result from the
increased base shear resulting from the post-yield stiffness of the structural response
causing additional rotation on the flexible base. In this case the displacement ductility
demand is found to be
Ap tAy
A, +A,
(Ala)
If the design displacement is limited by a code specified maximum drift, then
foundation flexibility will increase the yield displacement, bur not the design
displacement, and the design displacement ductility demand will be
Ap
7 Alb
wa (Ib)
In both cases, the design ductility demand will be less than for the equivalent rigid-
dase case, and hence the equivalent viscous damping will also generally be reduced.354
ley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Vacs g
& foundation
flexibility
a included
Displacement
(a) Structure (b) Force-Displacement
Fig.6.19 Influence of Foundation Flexibility on Design Displacement
However, as noted in Section 3.5.4(b) foundation deformation will generally also be
accompanied by additional damping, > resulting from hysteretic soil response, and
radiation damping, As shown in Section 3.5.4(b), this can be included in the DDBD
procedure by using a system equivalent viscous damping of
ihe tabs
(6.42)
Ae + As
$e.
where & is the structural damping associated with the structural displacement ductility
demand As/A, = (Ap - ‘Ar)/Ay. Limited experimental evidence? supports foundation
damping ratios in the range 0.05, for foundations responding without uplift, to 0.
foundations uplifting and reaching maximum overturning moment capacity.
6.5.2 Foundation Rotational Stiffness
Unless very massive foundation structures or support on piles with tension uplift
capacity are provided, some uplift on the tension edge on the foundation/soil interface
must be expected. This has a significant influence on the effective rotational stiffness of
the foundation, which must be included when estimates are made of the foundation-
induced displacement at the effective height. Consider the wall foundation shown in
Fig.6.20, The foundation has been sized to provide a static factor of safety against gravity
loads of 6, and the soil is represented by elasto-plastic response with a soil deformation at
yield of 25 mm (1 in). Under gravity loads the settlement is thus approximately 4 mm.
(0.16 in),
ers. comm. R. Paolucci“Chapter 6. Structural Wall Buiings 355
Rocking Displacement at He (9)
1. Gravity loads
2. Ma = PL+/6; 00
2 s. 3.M=2M5 0
jearing pressure
3 profiles at different 4M = 2.33Ms; 0 = 90
5 | stages of uplift 5.M = 2.48My; 0 = 278)
Fig.6.20 Foundation Compliance Effects for an Uplifting Spread Footing
The initial clastic rotational stiffness Kg can be determined by imposing a unit
rotation on the footing/soil interface resulting in
B,L;
D =kyly (6.43)
where Ay is the vertical subgrade modulus for the soil/foundation (kN/m, or kips/in’),
and By and Lr are the width and length of the footing/soil interface.
The foundation rotation due to a design base-shear force Vea for an clastic
foundation will thus be
9p =VaayelHe * hp)! Ko (6.48)
and the rotation-induced displacement at the effective height will be
Ap =O-(H. the )=V,(H, the) IK, (645)356 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
However, the above formulation applies for conditions where the footing remains in
contact with the soil along the full length of the footing, This is unlikely to be the case in
many spread footing designs, as it would require uneconomically large spread footings.
In fact, foundation rocking may be beneficial to structural response as additional
damping is provided, and the displacements resulting from foundation rocking may
reduce the structural displacement demand, and hence the damage potential.
Figure 6.20 includes bearing pressure profiles at different stages of uplift, numbered 2
to 5. At profile 2 the footing has zero stress at the tension edge, corresponding to the
limit for which Eqs.(6.43) to (6.45) apply. The peak bearing stress is twice the gravity load
bearing stress. At profile 3 the footing has uplifted over 50% of the length, and the
maximum beating stress is twice that for profile 2 (ic. four times gravity load bearing
stress). The moment required to develop this profile is twice that for profile 2, but the
foundation rotation has increased by a factor of 4. The effective stiffness is thus only
50% of that given by Eg, (6.43).
‘At profile 4, only 1/3" of the footing remains in contact with the soil, and the
maximum compression stress is 6 times the gravity load value — equivalent to the ultimate
bearing stress in this example. The moment is now 2.33 times that for profile 2, and the
rotation is 9 times larger, indicating an effective (secant) stiffness that is only 26% of the
fully elastic value. The footing can continue to rotate with small increase in moment
capacity, by plastic deformation of the soil (profile 5), with continual degradation of the
effective stiffness. For the conditions represented in this example, the shear
force/rocking, displacement relationship has been included in the plot of Fig.6.20, which
includes the bounding envelope resulting from P-A effects (see Section 3.6). Although
the peak lateral\force (at approximately 0.29P) seems large, it should be recalled that
structural walls geferally have substantially larger tributary areas for inertia force than for
gravity load. \
‘The recommendations for stability of cantilever walls (Section 6.1.3) suggested
matching the ultimate overturning moment capacity of the foundation to the input
corresponding to overstrength conditions at the base hinge. Assuming an overstrength
factor of § = 1.25, this would imply that conditions at the design lateral forces would
correspond almost exactly to profile 3 (uplift of 50% of the foundation length). ‘This limit
condition for design of spread footings is included in many design codes. It is apparent,
then, that the appropriate stiffness to use for estimating foundation compliance effects is
0.5Kawhere Kpis given by Eq.(6.43)
It is of interest to investigate the significance of foundation flexibility for the typical
example shown in Fig.6.20. Using standard DDBD procedures the effective height is
found to be 13.8 m (45.3ft). Flexural reinforcement expected yield strength is fye = 462
MPa (67 ksi).
= 0.00231 13, #1 BS ye 40.9mm (1.61 in)
3x18)
Foundation displacement: We assume that at design strength, the foundation uplift
corresponds to profile 3, as noted above. In this case we do not need information about
Structural yield: From Eq (6.5): A,Chapter 6, Structural Wall Buildings 357
the soil stiffness or the base shear force. The maximum bearing stress is (0.67py and hence
the maximum settlement is 0.67x25mm = 16.75 mm. The foundation rotation is hence
9 = 0.01675/6 =0.00279 radians, and the displacement at the effective height is
A, =6,(H, +h, )=0.00279(13.8+2.5)=45.5mm (1.79 in)
‘Thus the displacement at the effective height, resulting from foundation compliance, is
10% larger than the structural component of yield displacement. The influence on design
will be considerable.
6.6 CAPACITY DESIGN FOR CANTILEVER WALLS:
“The need for protection of locations and actions against unintended inelastic response
has been emphasised already in different parts of this text (e.g, Sections 3.9, 4.5 and 5.8)
In particular, the treatment of Section 3.9 used as an example 4 current, widely accepted
approach for determining the required distribution of flexural and shear strength up
cantilever walls to account for flexural overstrength at the wall-base plastic hinges, and
for dynamic amplification resulting from higher-mode contributions to response.
Te was noted in Section 5.8, in relation to capacity design of frames, that existing
methods for capacity protection did not adequately account for the influence of ductility
demand. Inelastic time-history analyses (THA) showed that when the intensity of
excitation (and hence the system ductility demand) was increased, the influence of higher-
mode effects in amplifying the envelopes of column flexure and shear also increased.
Simple design equations were presented to represent this effect.
Investigations into the response of cantilever wall structures!!! using ITHA has
indicated similar, though more pronounced trends. In this study, six walls, from 2 storeys
co 20 storeys were designed to a linear displacement spectrum with a cornet period at 4.0
sec, and corer displacement of 0.594 m (23.4 in) for 5% damping. This corresponds to a
PGA of 0.4g, and medium soil conditions.
‘The walls (see Fig.6.21) all had the same tributary floor mass of 60 tonnes, and gravity
load of 200 kN at cach level, and were designed in accordance with DDBD design
principles to achieve maximum drifts of about 0.02 at roof level. Wall lengths (fy), widths
8), reinforcement contents (g)) and bar sizes (dy) varied from wall to wall in order to
satisfy the design displacement criteria. Details are listed in Table 6.4, which also includes
the calculated plastic hinge length (Lp), the expected displacement and curvature ductility
demands (dg £4), the effective period at maximum displacement (Zz, approximately equal
to Ty[ddy ), and design base shear force and bending moment (Vase and Mpa)
Note that limiting the drift to 0.02 results in displacement ductility demands that are
Jess than typical code limits of ls = 5 in all but the two-storey wall. The six designs were
subjected to time-history analysis using a suite of five spectrum-compaible earthquake
records. These records were intensity-scaled tc 50%, 100%, 150% and 200% of the358 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
design intensity to investigate the sensitivity of the results to intensity, and hence to
displacement ductility demand.
2 WALL
T nen
‘Y i, 1 Yi
Fig.6.21Idealization of Cantilever walls in Capacity Design Studyi?21
Table 6.4° Wall Details for Capacity Design Study
ee b le dy Ip ws | py [Te (sec) Ve | Mp
@) | & (am) | (KN) | @cNm)
A 0.20 2.0 | 0.0046 14 0.58 64 20.6 1.2 242 | 1232
B_ | 020 | 25 [oooso[ 14 [oso [34 [126 [ 18 | 312 [2917
c 0.20 3.3 0.0162 20 149 19 6.0 2.6 446 8114
D 0.25 4.0 | 0.0172 28 2.22 13 27 3A 590_| 16222
E | 025 | 50 [oowi] 24 | 283 | 12 | 22 | 37 | 664 | 24372
[FT| 030756 [oor77 | a8 [352 [io | to [39830 38739
Averaged results from the ITHA are presented in Figs. 6.22 and 6.23 for moment
and shear envelopes respectively, and compared with values based on two different
current design approaches. The first design approach is the procedure represented in
Fig.3.28, and adopted in several design codes, and the second is a modal
superposition approach, also commonly specified in design codes, where the design
envelopes for shear and moment are determined from an clastic modal superposition
using the elastic acceleration spectrum, with the results then divided by the design
displacement ductility demand. In Figs.6.22 and 6.23, these alternative approaches are
identified as Cap.Des and SSRS/{« respectively. Note that the SSRS method of modalChapter 6. Structural Wall Buildings 359
combination was used in the latter case, but identical results would have been found from
the more rigorous CQC method since the modes were well-separated. In these figures
“IR” indicates the intensity factor applied to the standard spectrum-compatible records.
Thus “IR = 1.5” indicates 150% of the design intensity, and so on. The results, for both
ITHA and design methods include only the dynamic amplification, since material
overstrength was not included in the analyses, except for the proportion of overstrength
resulting from reinforcement strain-hardening.
Referring first to Fig.6.22, we see that the time-history analysis results indicate only
small increases in wall base bending moment with increasing intensity, as expected, since
the increase, once the nominal moment capacity has been reached is only the result of the
post-yield stiffness of the moment-curvature characteristic at the wall base. However, at
levels above the base, and particularly at wall mid-height, moments increase very
significantly with increasing intensity, especially for the eight- to twenty-storey walls, It is
apparent that both existing design procedures are non-conservative at the design
intensity, (TR = 1.0) and increasingly so at higher intensities. For the two- to cight-storey
walls, where the design displacement ductility exceeds 2 (see Table 6.4), the multi-modal
moment envelope is non-conservative even at 50% of the design intensity.
In Fig.6.23 it is again seen that the time-history shear force envelopes are strongly
influenced by seismic intensity, (and hence by ductility level), and that both the capacity
design and multi-modal design envelope are significantly non-conservative. For the ewo-,
four-, and cight-storey walls, the time-history base shear force at IR=1 is almost owice
the multi-modal value, with a slightly smaller discrepancy for the capacity design
cavelope, and for these three walls, the shear profiles at IR=0.5 exceed the design profile
at all heights. At intensity ratios of IR=2, base shear force is between 2.5 and 3.7 times
the multi-modal design envelope. For the taller walls the SSRS/H envelope exceeds the
capacity design envelope, and thus the discrepancy from the capacity design value is even
higher.
‘The discrepancies beaween the capacity design and time-history shear forces are more
problematic than the corresponding moment discrepancies of Fig.6.22, Although un-
intentional plastic hinging (which could be the consequence of designing to either the
Cap.Des or SSRS/I capacity moment envelopes of Fig.6.22) at levels above the base is
undesirable, some limited ductiliy demand should be sustainable without failure.
However, the consequences of the imposed shear demand exceeding the shear capacity,
by such large margins, could be catastrophic shear failure.
6.6.1 Modified Modal Superposition (MMS) for Design Forces in Cantilever
Walls
Examination of Fig.6.23 indicates thar at an intensity ratio of IR=0.5, where ductility
demand is low, or non-existent for all walls, the shape of the shear force envelope is well
predicted by the modal analysis procedure. This suggests that it might be possible to
predict the shear force and moment envelopes by simple modification of the modai
response spectrum (SSRS/j1) approach360 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Height (en)
Height (rm)
© 40000200 2 t000»—2000 000
Moment (kNen) Moment (Nm)
(a) Two-Storey Wall, () Four-Storey Wall
© 2000 4000 6000 6000 10000 © #000 e000 12000 18000, 20000
Moment (kim) ‘Momenc (km)
(© BightStorey Wall (@) Twelve-Storey Wall
Height (o»)
Height (om)
° 10000 20000 s0000 @ 1800 20000 080 40000
Moment (kNNm) Moment (kNm)
(6) Sixteen-Storey Wall (0) Twenty Storey Wall
Fig.6.22 Comparison of Capacity Design Moment Envelopes with Results of
‘Time-History Analyses for Different Seismic Intensity Ratios!”21Chapter 6, Structural Wall Buildings
361
. 2
os] sofss fw
at Jsj0 is |20 oa)
& €
] + susod
cap they,
oT a ooo
© 100 200 200 40900 0200 400600. at 1000
Shear Force (KN) Shear Force kN)
(@) TworStorey Wall (6) FourStorey Wall
0
=
En
a
w4 “LO
20 swt
cpoe4 4 u tL
a 409 e60 1200. 1600 2009 © 00 1000 1500. 2000 2500
shear Force (kN) Shear Force (kN)
(©) Bight Storey Wall
Height (n)
° 000
2000
Shear Force (kN)
(©) Sintcen-Storey Wall
000
Height (m)
(8) Twolve-Storey Wall
2000 000
Shear Force (KN)
(6 Twenty Storey Wall
4000
Fig.6.23 Comparison of Capacity Design Shear Force Envelopes with Results of
‘Time-History Analyses for Different Seismic Intensity Ratios!?21362 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
‘A basic and simple modification to the modal superposition method is available by
recognizing that ductility primarily acts to limit first-mode response, but has
comparatively lite effect in modifying the response in higher modes. If this were to in
fact be the case, then first-mode response would be independent of intensity, provided
that the intensity was sufficient to develop the base moment capacity, while higher modes
would be directly proportional to intensity. This approach is very similar to that proposed
by Bibl and Keintzell™ as a means for predicting shear demand at the base of cantilever
walls.
This modified modal superposition approach is clearly an approximation to response.
Although the first-mode inelastic shape is very similar to the elastic shape, and hence the
approximation should be reasonaily valid for the first mode, it is clear that the higher
‘modes will be modified to some extent by the first-mode ductility, since a basic feature of
the modified higher modes will be that, when acting together with ductility in the first
mode, they cannot increase the base moment demand, which will be anchored by the
moment capacity of the base plastic hinge. The approach suggested below extends the
basic method of Bibl and Kreintzel for shear forces to the full height of the wall, and also
provides a method for determining the appropriate capacity-desiga moment envelope.
Modifications to this approach are discussed in relation to dual wall/frame structures in
Chapter 7, and to bridges in Chapter 10. A brief discussion of possible further
improvements in included in Section 6.6.1.(C).
(a) Shear Force Profiles: To investigate the appropriateness of a simple approach based
fon the above arguments, shear force profiles were calculated based on the following
aesumpéons
* First-mode shear force was equal to the shear profile corresponding to
development of the base moment capacity, using the displacement-based design
force vector. However, for low seismic intensity, where plastic hinging was not
anticipated in the wall, simple elastic first mode response, in accordance with the
elastic response spectrum was assumed.
‘© Higher-mode response was based on elastic response to the acceleration spectrum
appropriate to the level of seismic intensity assumed, using the elastic higher-mode
periods. Force-reduction factors were not applied
‘© The basic equation to determine the shear profile was thus:
Vans; = Wing +Vies + Viet) (6.46)
where Vays, is the shear at level 4 Vip, is the lesser of elastic first mode, ot ductile
(DBD value) first-mode response at level 4 and Vagi, and V3e; etc are the elastic modal
shears at level 7 for modes 2, 3 etc. Predictions for shear force profiles based on this
equation are included in Fig.6.25.
(b) Moment profiles: A simple modal combination, similar to that of Eq.(6.46), but
multiplied by a factor of 1.1, over the top half of the wall, with a linear profile from mid-Chapter 6. Structural Wall Buildings 363
height to the moment capacity at the base of the wall was found to provide best results
for moment profiles (see Fig.6.24). ‘The combination equation over the top half of the
wall is thus:
Myysy = 1M Miny + Mie, + Moe, Fo)? (647)
where Mus; is the moment at level 4 Mypy is the lesser of the elastic first mode
moment and the ductile design moment, and Mx; and Mgej ete ate the clastic modal
moments at level for modes 2, 3 ete:
Iris seen that the MMS approach provides a good representation of the time-history
moment profiles in Fig.6.24 at the design intensity (IR=1.0), for all walls. ‘There is a
tendency for the MMS predictions to be slightly unconservative for the shorter walls, and
slightly conservative for the taller walls, though the discrepancies are generally small. ‘The
change in shape of the moment profiles with increasing intensity is also well represented
by the MMS predictions. As discussed above, slight unconservatism in the moment
profiles is acceptable, as it implies only limited ductility demand, and there is a case for
deleting the 1.1 factor in Eq.6.47.
Similar behaviour is apparent for the shear force comparisons of Fig.6.25. At the
design intensity the agreement between the MMS and THA profiles is extremely close
for the four- to twenty-storey walls, and is adequate, though a little unconservative for the
nwo-storey wall. Similar conclusions apply at different intensity levels, though the MMS.
approach becomes increasingly conservative for the taller walls at high intensity ratios.
(c) Effective Modal Superposition: \s noted in Chapter 10 with relation to capacity
design of bridges, a simple and philosophically attractive (from a DDBD viewpoint)
further modification to the MMS approach is to carry out the modal analyses using
effective stiffness of members at maximum displacement response, as is used in the
design process. Thus, for a cantilever wall, the stiffness of the first storey would be
reduced in proportion to the displacement ductility demand when defining the structure
for modal analysis. With this modification, the approach is identical to that presented in
she previous section.
6.6.2 Simplified Capacity Design for Cantilever Walls.
In many cases the additional analytical effort required to carry out modal analysis of
the designed wall structure to determine the capacity design distribution of moments and
shears will be unwarranted, and a simpler, conservative approach may be preferred. The
following approach, based on the data presented herein, is suggested.
The results for both moment and shear force envelopes indicate that dynamic
amplification increases as the intensity ratio increases. ‘This would indicate that
displacement ductility demand should be included in the design equation. Also, it would
appear obvious that the number of storeys, which has been used in the past as a key364
Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Shomen: Ne)
(oy Tore Soy Wa ROS
Haig cn
reas
omen (tomy
(a) Siven Soy Wa, IR
Hehe
Hiden ey
Mlomen qe
(0) Toe Stoney Wa IRAN
‘omen fn
(6) eee Sey Wal, IRL
somes (my
(o)Soncen Sey Wa, RL
Moment Ne)
(0) owen storey Waly IRL
an =
t &
nyse sae se as Dae an" aan tan
aomer ih) ‘Moment Ne)
(ey Twelve ioe Wa, RLS (oy Teen Say Wa
as
Heighe (oy
‘omen: Gm)
(Toe Sty Wa R20
Megh o
‘omen ee)
(9 Been ery Wa, PR=ZO
see se
Fig.6.24 Comparison of Modified Modal Superposition (MMS) Moment
Envelopes with ITHA results, for Different Seismic IntensitiesChapter 6, Structural Wall Buildings 365
3
2
‘ear Fore GN) ‘Shear Fore G2 ‘Shea Fore en
(on Tt Sorey Wal, ROS (op Sistensorey Wa R= (9 Teen Storey Waly IR=0S
:
(oy Taahve Sey Wa
(ey Twenty Soney Way R10
(oy twalee Song WN IR=20 (o Sacer Sony Wal IR=20 (0) Tosh Sea DP rea
Fig.6.25 Comparison of Modified Modal Superposition (MMS) Shear Force
Envelopes with ITHA Results, for Different Seismic Intensities366 Priestley, Calvi and Kowalsky.Displacement-Based Seismic Design of Siractures
parameter for capacity design (see Eq.(3.62), e.g.) should be less significant than the
fundamental elastic period J; of the wall. ‘This leads co the following recommendations:
(4) Moment Capacity-Design Envelope: A bilinear envelope is defined by the
overstrength base moment capacity gMg, the mid-height overstrength moment Mp5,
and zero moment at the wall top, as illustrated in Fig.6.26(a) for a four-storey wall. The
overstrength base moment is determined from section and reinforcement properties, as
suggested in Section 4.5, The mid-height moment is related to the overstrength base
moment by the equation:
Mosin=Cip O'My, where Cy ossoorst{ £-1]z08 (6.48)
4 TI
Bo
Tension
4m cin
= Capacity z envelope
2 2 ‘
= q
L
Ye
>
$M; ———4 Vane= 0O/Vase —al
(a) Moment Capacity Envelope _(b) Shear Force Capacity Envelope
Fig. 6.26 Simplified Capacity Design Envelopes for Cantilever Walls
Note that U/@ is the effective displacement ductility factor at overstrength, and thar
tension shift effects should be considered when terminating flexural reinforcement
Tension shift, resulting from inclined flexure/shear (diagonal tension) cracking results in
flexural reinforcement stress at a given level being related to the moment at a level closer
to the wall base. In effect, this “shifts” the design moment profile upwards, as suggested
by the upper dashed line in Fig,6.26(a). The tension shift depends on wall length and the
amount of transverse reinforcement provided, but it is reasonably conservative to assume
a tension shift equal to y/2, where fy is the wall length. A complete discussion of
tension shift is available in (P1]
(®) Shear Force Capacity-Design Envelope: The shear force capacity envelope is
defined by a straight line between the base and top of the wall, as indicated in Fig.6.26(b).Chapter 6. Structural Wall Buil
ings 367
‘The capacity-design base sheat force is related to the DDBD base shear force by:
Virose =P OV pose (6.49)
where @
Or and Cy =0.067+0.4(T, —0.5)< 1.15 (6.50)
‘The design shear force at the top of the wall, V’, is related to the shear at the bottom
of the wall by:
Vr = CV ease
where C,=0.9-0.37, 203 (6.51)
2qs.(6.48), (6-50) and (6.51), Tj is the elastic fundamentai period.
Predictions for the ratio of wall moment at mid-height to base moment, and dynamic
amplification factor for base shear force are compared with values obtained in the ITHA
for different elastic periods and ductility levels in Fig.6.27
2%
1 YT=198 zs
z
0.8 34
; a
206 23
§ 2
E04 22
Eo.
2 é
02 a
&
ot T T T 1 oF T T T 1
° 2 4 6 8 0 2 4 6 8
Displacement Ductility Demand, It Displacement Ductility Demand,
(a) Ratio of Midheight to Base Moment _(b) Base Shear Dynamic Amplification
Fig.6.27 Comparison of Capacity Design Equations (6.48) and (6.50) with Time
History Results for Different Elastic Periods and Ductility Levels
In Fig.6.27, the results from the ITHA are shown by solid data points, with the same
symbol used for all different ductility levels (ie. different seismic intensities) for a given
wall. Predictions by the equations are shown as continuous lines. Agreement is good for
both mid-height moment ratio and base shear force. The slightly unconservative nature
of the wall mid-height moment ratio prediction has been deliberately imposed since
minor inelastic response at levels above the base is acceptable. However, it should be368 Priestley, Calvi and Kowalsky. Displacement-Rased Seismic Design of Structures
noted that since nominal flexural strength will be matched to the capacity envelope, and
since inelastic response occurs at moments lower than the nominal moment of the
bilinear moment-curvature approximation (see Fig.4.6(a), €4), the inelastic response may
in some cases be significant.
(6) Strength Redaction Factors for Capacity Design: When determining the required
amount of sansverse reinforcement for shear strength, a strength reduction factor of v
= 0.85 should be used, together with conservative estimates of material strength, as
discussed in Section 4.5. However, with flexural strength this may not be practicable, nor
necessary. Since moment demand may reduce only slowly up the wall, designing for
conservative material strengths together with a flexural strength reduction factor can
mean that flexural reinforcement content is required to increase at levels above the base,
particularly when it is remembered thar axial load, which contributes to flexural strength
of walls will decrease with height. Recognizing, again, thar the consequences of minor
inelastic flexural action at levels above the base are acceptable, we recommend that
flexural reinforcement areas at levels above the base be determined using the same
expected material strengths used to design the wall base, without inclusion of a flexural
strength reduction factor. This is of course reasonable, as the flexural reinforcement at
the base is likely to extend, with uniform strength for a considerable height above the
base (note that use of short starter bars with lapping of flexural reinforcement at the wall
base is undesirable, as the strength of lap-splices ends to degrade under repeated load
reversals, and the plastic hinge length is condensed below levels implied by Eq,(6.7)).
(d) Overstrength factors for Capacity Design: A consequence of the argument
presented in the previous section is that the flexural overstrength factor adopted in
Eq,(6.48) should only inchade the component resulting from strain-hardening, and not
from excess yield strength. Since strain-hardening will normally be included in the
DDBD process for determining required flexural reinforcement at the wali base, this
implies thac @° = 1.0 for flexural design. If strain-hardening is ignored in determining
required base flexural reinforcement content, a value of g = 1.2 should be adopted, as
implied by Section 4.5.2
For shear design, the value of @ should include allowance for material oversirength,
strain-hardening, and excess flexural reinforcement over that required to provide the
design strength, if provided, and should normally be determined by moment-curvature
analysis. If not, the values recommended in Section 4.5.2 should be adopted.
(©) Design Example 6.2: Capacity Design of a Wall Building: ‘Vhe capacity design
moments and shear forces for the walls of Design Example 6.1 (Section 6.4.6) are to be
determined using the simplified approach of Section 6.6.2. Note that moments and shear’
for the walls corresponding to the distributed base shear force are already included in
Table 6.3.Chapter 6, Seructural Wall Buildings 369
Capacity Moments: Using, the recommendations in Section 6.6.2(d) the over-
strength factor to be used in Eq,(6.48) is taken as $? = 1.0, From Section 6.4.7 the design
system ductility was found to be pays = 5.0. Equation (6.48) requires the initial (elastic)
period T; to be calculated. This can be estimated from the effective period T, using the
relationship:
(6.52)
where ris the ratio of post-yield to elastic stiffness (refer Fig.3.1, e.g.). Taking a typical
value of r= 0.05, and the effective period of T, = 1.97 sec, the initial period is found to
be T= 0.975 sec.
“The mid-height moments are found from Eq,(6.48) as follows:
Coefticient Cy. Cp =0.4+0.075T,(4t/ 1-1) = 0.4 + 0.075 0.975x4 = 0.693
8m Wa Mocy =C,r0° My =0.693X1X25100=17400kNm (153,000 kip in)
4m Wall: Moy, =0.693X16300=11300kNm (100,000 kip.in)
‘The corresponding design overstrength moments are listed as Me in Table 6.5,
together with the moments Mj corresponding to the distributed base shear force from
Table 6.3. Note that when designing the reinforcement content for each wall, tension
shift should be applied to the moments of Table 6.5 in accordance with the
recommendations of Section 6.6.2(2).
Capacity Shear Forces: Moment-curvature analyses using the maximum feasible
yield strength of the reinforcement of fy’ = 1.3f = 546MPa (79.2 ksi) show that at the
design curvatures (see Example 6.1) the overstrength factors for both walls are g? = 1.09.
Wall Base: From Eq.(6.50): C,, = 0.067 + 0.4(T, ~0.5) = 0.067 + 0.4x0.475 = 0.257
and, =1+(u/ Cy, =1+(5/1.09)0.257 = 2.18
Hence from Eq.(6.49):
Van, = 0° OV py. p = 1.09% 2.18% 2040 = 4847KN (1089 kips)
COV gaury =1.09X2.18X1316 = 3127KN (703 kips)
Wall Top: The wall top shear force is related to the wall base shear force by Eq.(6.51),
where C, =0.9-0.37, 20.3=0.9-0.3x0.975 = 0.608. Thus:
8m Wall: Vi =C,Vg,, = 0.608 4847 = 2944kN (662 kips)
4m Wall: V2 = CV, =0.608X3127=1901KN (427 kips)
“The corresponding design shear forces are listed as Ve in Table 6.5, together with the
shears, Vj, from distribution of the design base shear force from Table 6.3.370 Priestley, Calvi and Kowalsky. Displacement-Based Seismic Design of Structures
Table 6.5 Capacity Moment Shears and Moments for Walls of
Examples 6.1 and 6.2.
Floor | Height | Wall? | Wall1 | Wall1 | Wall1 | Wall2 | Wall2 | Wall2 | Wall2
(im) Vi ve M, Me Vi ve M Me
NY | GND) | Gedy | ce | GND | GND | CN | GN
0 0
168 | 606 | 2940 ce)
14 i100 | 3260 | 1700/5709 | 724 [210s [21 | 3770
112 | 1490 [3580 | 4780 [11300 | 974 | 2310 | 3150 | 7540
a4 | 1770 | 3900 | 8960 [ 17400 | 1152 | 2510 | 5880 | 11300
5.6 | 1950 | 42i0 | 13920 | 19900 [ 1264 | 2720 | 9100 | 12900
2.8 | 2040 [4530 | 19390 | 22500 | 1316 | 2920 | 12640 | 14600
Jo] }ro} es] a} ox}
0 2040 [4850 | 25100 [25100 [ 1316 [3130 [ 16330 | 16300
PRECAST PRESTRESSED WALLS
In Section 5.11.3 the concept was introduced of providing flexural strength and energy
dissipation to precast concrete frames by prestressing beams through the columns with
unbonded post-tensioned tendons, supplemented by bonded mild-steel reinforcement
grouted into ducts passing through the column and into the beams on either side (so-
called “hybrid” design). It is obvious that the concept of unbonded prestressing can also
be applied to precast wall buildings. Fig. 6.28 shows two possible applications. In the first
(Fig.6.28(@)), wall elements are stacked vertically aad post-tensioned to a foundation, This,
essentially creates a structural system that behaves in the same way as a prestressed beam-
to-column connection. A single crack can be expected to form at the critical section ar
the wall base, and will have non-linear clastic force-deformation characteristics, Note,
however, that the footing must be of sufficient size and weight to ensure that rocking
does not develop at the soil/footing interface. Note also thac the gravity weight
contributes to the flexural resistance of the wall base in exactly the same way as does the
prestressing,
Figure 6.28(a) also shows additional mild steel bars running through the wall footing
interface. ‘The normal way this would be achieved would be with bars cast into, and
protruding below the lowest precast wall element, grouted into preformed holes ia the
foundation. These mild-steel bars are expected to yield in tension and compression as the
base crack forms, in much the same fashion as conventional reinforcement. Since the gap
opening at the wall base can be large, it is normal to de-bond the mild stecl bars for some
Jength into the precast panel, to ensure strains at maximum displacement response are
kept to acceptable limits — normally less than 3%.
Provided that the combined axial force at the wall base resulting from gravity weight
and prestressing exceeds the compression yield force of all the mild steel bars crossing
the base interface, the residual displacements at zero lateral force will always be zero. In
this case the mild steel bars provide additional strength and damping to the bilinear clastic
response, resulting in the “flag-shaped” hysteresis loops of Figs.4.33 and 5.36, Wall
designs using this concept have been testedlH with excellent results.