You are on page 1of 335

CUREe-Kajirna Research Project

Final Project Report

Design Guidelines for Ductility and Drift Limits

By

Mr. Nobumasa Tanaka


Dr. Norio Inoue
Mr. Takaharu Fukuda
Mr. Hitoshi Hatamoto
Mr. Yoshio Sunasaka
Mr. Satoshi Ohrui
Mr. Tetsuya Tsujimoto

Prof. Vitelmo V. Bertero


Prof. Gary C. Hart
Prof. James C. Anderson
Prof. Helmut Krawinkler
Prof. Jack P. Moehle
Mr. Eduardo Miranda
Mr. Aladdin Nassar
Mr. Mohsen Rahnama
Mr. Chukwuma G. Ekwueme
Mr. Thomas A. Sabol
Mr. Xiaoxuan Qi

................

\
\
\

...........

............

......,

.,......

_,..,.J'
Report No. CK 92-03B
February 1992

California Universities for Research


in Earthquake Engineering (CUREe)

Kajirna Corporation

CUREe (California Universities for Research in Earthquake Engineering)

California Institute of Technology


Stanford University
University of California, Berkeley
University of California, Davis
University of California, Irvine
University of California, Los Angel~s
University of California, San Diego
University of Southern California

Kajima Corporation

Kajima Institute of Construction Technology


Information Processing Center
Structural Department, Architectural Design Division
Civil Engineering Design Division
Kobori Research Complex

'

', l
t

ABSTRAcr AND ACKNOWLEDGEMENTS

This report summarizes each of the studies that have been conducted in California as
a part of the CUREe-Kajima Research Project #5, entitled .. Design Guidelines for
Ductility and Drift Limits." This research project has been supponed by a grant
provided by the Kajima Corporation
Universities for Research in Earthquake

and administered

Engineering).

by CUREe

(California

This financial support is

gratefully acknowledged.
The report consists of seven chapters. The first six chapters summarize the six different
srudies that have been conducted according to the agreed team research project plan.

These srudies are described in

~etail

in the seven CUREe-Kajima reports given below.

REPORTS

1.

..

Bertero, V. V ., Anderson, J.C., Krawinkler, H., Miranda, E., "Design Guidelines


for Ductility and Drift Limits: Review of State-of-the-Practice and of-the-Art on
Ductility and Drift-based Earthquake Resistant Design of Buildings," July, 1991.

2.

Krawinlder, H., Nassar, A., and Rahnarna, M., "Evaluation of Damage Potential
of Recorded Ground Motions, .. June, 1991.

3.

Bertero, V. V., and Miranda, E., "Evaluation of Damage Potential of Recorded

Ground :Motions and its Implications for Design of Structures, July, 1991.

4.

Miranda, E., and Benero, V. V., "Evaluation of Seismic Performance of a TenStory RC Building," July, 1991.

5.

Anderson,

J.C., M"U"anda,

E., and Bertero, V.V., "Evaluation of Seismic

Perfonnance of a Thirty-Story RC Building, July, 1991.

6.

Hart, G., and Ekwueme, C. G., Japanese Concrete Frame Building Response,"
July, 1991.

7.

Qi, X., and Moehle, J.P., "Displacement Design Approach for Reinforced
Concrete Structures Subjected to Earthquakes," January, 1991.

Chapter 7, after a brief review of the studies reported in the above seven reports,
(summarized in the first six chapters), presents guidelines for the development of a
reliable method for estimating the values of response reduction factor R and discusses
how these values could be used to improve present U.S. and Japanese code procedures
for earthquake resistant design.

This report summarizes only the work done by researchers of the C'UREe -team. The
valuable contributions of the Kajima team to this joint research project are recognized
and gratefully aclalowledged.

'

TABLE OF CONTENTS

Chapter
1

Title
Review of State-of-the-Practice and

1.1-1.17

-of-the-Art on Ductility and Drift-based


Earthquake-Resistant

Design

by Vitelmo V. Bertero, James C. Anderson, Helmut


Krawinkler, and Eduardo Miranda
2

Evaluation of Damage Potential of Recorded

2.1- 2.19

Ground Motions

b'y Helmut Krawinkler, Aladdin Nassar, and Mohsen


Rahnama.
3

Evaluation of Damage Potential of Recorded

3.1- 3.19

Ground Motions and its Implications for Design


of Structures

by Vitelmo V. Bertero and Eduardo .Miranda

U.S. Concrete Frame -Building Response

4.1- 4.50

by James C. Anderson, Vitelmo, V. Bertero,


and Eduardo Miranda

Japanese Concrete Frame Building Response

5.1- 5.49

by Gary C.Hart and C.G. Ekwueme

Member Details and Response Reduction

6.1 .. 6.22

by Jack P. Moehle
7

Summary, Conclusions, and Implications


for Design
by Helmut Krawinkler and Vitelmo V. Bertero

7.1-7.9

DESIGN GUIDELINES FOR DUCfll.ITY AND DRIFT LIMITS:


REVIEW OF STATE-OF-TilE-PRACTICE AND OF-THE-ART ON
DUCTILITY AND DRIFT-BASED EAR1HQUAKE-RESISTANT
DESIGN OF BUILDINGS

A REPORT ON TASK 1 OF THE CUREe-KAJIMA RESEARCH PROJECT ON


DESIGN GUIDELINES FOR DUCfll.ITY AND DRIFT LIMITS

by
Vitelmo V. Bertero

James C. Anderson
Helmut Krawinkler
(

Eduardo Miranda

and
The CUREe and The Kajima Research Teams
/

A CUREe-KAJIMA RESEARCH REPORT


July 1991

ACKNOWLEDGEMENTS
This report summarizes the results obtained in the studies that have been conducted under
task 1 of a research project on topic 5 entitled "Design Guidelines For Ductility And Drift
Limits." This particular project is part of a program of research which is supported by a
grant given by Kajima Corporation and administered by CUREe, the California Universities
for Research in Earthquake Engineering. The financial support from Kajima is gratefully
acknowledged.

The main authors of this research paper wish to thank Mr. Nobumasa Tanaka, Group
Manager of the Kajima Research Team, and the members of this team for their valuable
comments. Thanks are extended to Hatem Goucha and Brad Young for their assistance
during the final preparation of this report.

TABLE OF CONTENTS
1. IN1'RODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1. 1 INTRODUCTORY REMARKS
.............................
1. 2 OBJECTIVES
............ ...............................
1. 3 SCOPE ................................................

-1-1-3-4-

2. NEEDS FOR DUCTILITY AND CONTROL OF DRIFT AND TIIEIR


PROPER USE IN ESTABLISHING RELIABLE EQRD CRITERIA ...... -52. 1 NEEDS FOR DUCTILITY
................................. -52. 1. 1 General Remarks ................................... -52. 1. 1 (a) Needs for Recognizing the Differences Between
Deformability, Ductility and Ductility Ratio . . . . . . . . . . . -62. 1. 1 (b) Advantages of Providing Structural Components and
Their Connections
with the Largest
Ductility
Economically Feasible .......................... -62. 1. 1 (c) Quantification of the Ductility Ratio ............. -82. 1. 1 (d) Concluding Remarks ........................ -92. 2 NEEDS FOR CONTROLLING INTERSTORY DRIFT INDEX
(IDI) .................................................. -92. 2. 1 General Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -92. 2. 2 Drift Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -102. 2. 3 The Need for Drift Design ........................... -11Table ................................................ -15Figures ............................................... -173. STATE-OF-THE-PRACTICE AND STATE-OF-THE-ART
OF EQRD OF RC STRUCTURES ...............................
3. 1 PROBLEMS IN DESIGN AND CONSTRUCTION OF EQSTRUCTURES ............ .............
RESISTANT
3. 2 STATE-OF-THE-PRACTICE
..............................
3. 2. 1 Estimation of Demands in Present Seismic Codes ..........
3. 2. 1 (a) Strength .................................
3. 2. 1 (b) Stiffness and Drift .........................
3. 2. 1 (c) P-A Effects ..............................
3. 2. 2 Estimation of Supplies in Present Seismic Codes ..........
3. 2. 2 (a) Strength. . ...............................
3. 2. 2 (b) Stiffness, Deformation and Stability Capacities ....
3.3 STATE-OF-THE-ART IN DUCTILITY AND STORY DRIFT-BASED
EQRD ...............................................
3. 3. 1 General Introductory Remarks Regarding Importance of J..L 6
and IDI in Preliminary Design .........................
3. 3. 1 (a) Examples ................................
3. 3. 2 (b) Concluding Remarks .......................
3. 4 STATE-OF-THE-ART IN USING DUCTILITY RATIO J..L 6 IN
PRELIMINARY EQRD ..................................

-21-21-

-22-22-22-23-24-24-24-25-26-26-27-32-32-

3. 4. I Use of J. 6 in Establishing the Design EQs ................


3. 4. 2 Use of J. 6 in Preliminary Design .......................
3. 4. 3 Concluding Remarks ...............................
3. 5 STATE-OF-THE-ART IN USING INTERSTORY DRIFT INDEX
(IDI) IN PRELIMINARY EQRD ...........................
3. 5. I Control of IDI at Serviceability Level ...................
3. 5. 2 Control of IDI at the Ultimate or Safety Limit State .......
3. 5. 3 Choice of Member Stiffness for Drift and P-.i Analyses ......
3. 5. 4 Recommended Practical Methods for Designing Considering
IDI. . ...........................................
3. 5. 5 Need to Consider the Effect of Multicomponent Seismic
Excitation and Direction in Estimating Structural Response
(IDI) ...... .- ....................................
Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4. STATE-OF-THE-PRACTICE:

REVIEW OF CODES ...................


4. 1 COMPARISON OF THE BASIC SLEDRS: UNITED STATES' UBC
AND JAPAN'S BUILDING STANDARD LAW (BSL) ...........
4. I. 1 Shape of the SLEDRS ..............................
4. I. 2 SLEDRS Values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2 COMPARISON OF THE DESIGN SPECTRA ..................
4. 2. I Japan's BSL. . .....................................
4. 2. 1 (a) Service-Level Design Earthquake ..............
4. 2. 1 (b) Ultimate or Safety Level Design Earthquake. . ....
4. 2. 1 (c) Comparison of Japan's BSL and UBC Design
Earthquakes: Service Limit State Level . . . . . . . . . . . .
4. 3
OBSERVATIONS REGARDING DERIVATION OF SIDRS
DIRECTLY FROM SLEDRS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4. 3. I Tri-Services Manual Approach .......................
METHOD 1:
ELASTIC ANALYSIS
4. 3. I (a)
PROCEDURE ...............................
METHOD 2:
CAPACITY SPECTRUM
4. 3. I (b)
METHOD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4. 3. 2 Comparative Designs of Buildings Using U.S. and Japanese
Codes ...........................................
4. 4
NEW ZEALAND CODE OF PRACTICE FOR GENERAL
STRUCTURAL
DESIGN AND DESIGN LOADING FOR
BUILDINGS (NEW ZEALAND STANDARD NZS 4203: I984) ....
4. 4. 1 Material Code Format ..............................
4. 4. 2 Material Strength ..................................
4. 4. 3 Strength Reduction Factor ...........................
4. 4. 4 :Load Factor U ....................................
4. 4. 5 Special Observations or Aspects .......................
4. 4. 5 (a) Earthquake Provisions .......................
4. 4. 5 (b) Method of Analysis ........................

-32-33-34-35-35-36-36-38-39-4I-45-45-45-45-45-45-46-47-48-52-53-54-54-55-56-56-56-56-56-57-57-57-

4.4.6 Equivalent Static Force Analysis:

4. 5
4.6

4. 7

4. 8

Total Horizontal Seismic


Force or Shear ....................................
COMPARISON OF REQUIRED DESIGN SEISMIC FORCES AT
SAFETY LEVEL BY NZS WITH UBC AND JAPAN'S BSL ......
COMPARISON OF REQUIRED LIMITS FOR THE INTER-STORY
DRIFT BY NZS, UBC AND BSL ...........................
4. 6. 1 NZS Requirements ................................
4. 6. 1 (a) Calculations of Deformations. . ...............
4. 6. 1 (b) Building Separation ........................
4. 6. 2 UBC Requirements ................................
4. 6. 3 Japan's BSL Requirements. . .........................
MEXICO: 1987 TECHNICAL NORMS (STANDARDS) FOR THE
FEDERAL DISTRICT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4. 7. 1 Mexico Required Strength at Ultimate (Safety) Limit State ..
4. 7. 2 Mexico Requirements for lnterstory Drift . . . . . . . . . . . . . . . .
DUCTILITY AND DRIFT CONSIDERATIONS IN EUROPEAN
SEISMIC CODES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4. 8. 1 Reference Documents ..............................
4. 8. 2 Summary of Relevant Aspects of Design Procedure ........
4. 8. 2 (a) Seismic Input .............................
4. 8. 2 (b) Design Spectrum ...........................
4. 8. 2 (c) Behavior Factors. . .........................
4. 8. 2 (d) Drift Considerations. . ......................
4. 8. 2 (e) Other Relevant Considerations ................
4. 8. 3 Evaluation of Ductility Considerations ..................
4. 8. 4 Evaluation of Drift Considerations : ....................
4. 8. 5 Summary ........................................
Tables ................................................
Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5. SUMMARY OF STATE-OF-THE-PRACTICE ON DUCTILITY


AND DRIFT-BASED EARTHQUAKE-RESISTANT DESIGN ..........
5. 1 GENERAL REMARKS ...................................
5.2 CODE SPECIFIED SLEDRS ...............................
5. 2. 1 Sites with Firm Soils (Soil Type 1). . ...................
5. 2. 2 Sites with Soft Soils (Soil Type 3). . ....................
5. 3 USE OF J.L 6 TO REDUCE SLEDRS TO SIDRS .................
5. 3. 1 Firm Soil (Soil Type 1). . ...........................
5. 3. 2 Sites with Soft Soils (Soil Type 3) .....................
5. 4 STATE-OF-THE-PRACTICE ON THE USE OF IDI LIMITATIONS
IN EQRD
............................................
5. 4. 1 Minimum Lateral Stiffness and Acceptable Limits on IDI at
Serviceability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5. 4. 1 (a) Short T .................................
5. 4. 1 (b) Long T .................................

-57-59-

-61-61-61-62-62-63-

-64-65-65-

-66-66-66-66-67-68-68-69-69-71-71-

-73-75-

-87-87-

-88-88-89-89-89-90-90-91-91-92-

5. 4. 2 Maximum Acceptable IDI at Ultimate Limit States (Collapse) .

-92Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -96

6. RESEARCH, DEVELOPMENT, AND EDUCATIONAL NEEDS .......... -976. 1 GENERAL REMARKS REGARDING THE NEED FOR IDEAL
SOLUTION ..................................
-976. 2 RESEARCH AND DEVELOPMENT NEEDS TO IMPROVE THE
ESTABLISHMENT OF SLEDRS . . . . . . . . . . . . . . . . . . . . . . . . . . . -996. 2. 1 SLEDRS for Service Level EQs. . . . . . . . . . . . . . . . . . . . . . . -996. 2. 2 SLEDRS for Survivability .......................... -1006. 3 RESEARCH AND DEVELOPMENT NEEDS TO IMPROVE THE
ESTABLISHMENT OF SIDRS ............................ -1006. 3. 1 SIDRS for Strength, Cy ........................... . -1016. 3. 1 (a) Code Procedures to Determine SIDRS for Cy ... . -1026. 3. 2 Implications of Recent Research Results with Respect to
-104Rationale for R Code Values ............
6. 3. 2 (a) The Case of Rock and Firm Soil Sites ........ . -1066. 3. 2 (b) Case of Very Soft Soil Sites ............... . -1066. 3. 2 (c) Recommendations for Improving Code SIDRS .. . -1086. 3. 2 (d) SIDRS for Lateral Displacement and IDI ...... . -109Table
-111Figures ............................................. . -113o

7. SUMMARY, CONCLUSIONS, AND RECOMMENDATIONS ............


7. 1 SUMMARY ................
7. 2 CONCLUSIONS ........................................
7.3
RECOMMENDATIONS .................................
7. 3. 1 Recommendations for Improving Code SIDRS for Strength,
Cy .............................................
. 7. 3. 2
Recommendations for Improving SDIRS for Lateral
Displacement and IDI. .............................

-131-131-132-144-

8. REFERENCES

-146-

-144-145-

1. INTRODUCfiON

1.1 INTRODUCfORY

REMARKS

One of the most promising approaches for developing efficient methods for improving
earthquake-resistant design (EQRD) is predicting the response of structures to earthquake
(EQ) ground motions through the use of an energy balance equation. The energy balance
equation for a general oscillatory system subjected to base acceleration can be written as:
~ =

where

Es

+ Eo

(1. 1)

= total input energy

Es
Eo

= energy stored in the structure


= energy dissipated from the structure

The stored energy and dissipated energy can further be defined as


Es=~+~

(1. 2)

and

Eo
where

= E,

+~

(1. 3)

~ =

elastic strain energy

~ =

kinetic energy

E, = energy dissipated through viscous damping


~

= energy dissipated through hysteretic, inelastic deformations

For a linear elastic system the

can be expressed as follows


(F_J(v)

(1. 4)

2
where F. is the restoring elastic force and vis the relative displacement of the reactive mass,
equal to F/K where K is the elastic stiffness.

From Eq. 1.4 it is clear that for any given linear elastic perfectly plastic system the maximum
~

that can be stored is reached when F. reaches the elastic limit or yielding force Fy of the

system. Therefore to increase the

, and thereby the

Es , it is necessary either to increase

Fy and/or to increase the yield displacement vr by decreasing K. Because the increase of

-2Fy usually demands large amounts of material, and therefore an increase in the weight and
cost of the structure, it would seem logical to try to increase vy as much as possible by
decreasing K. However, because of the need to have a serviceable structure and to control
damage of nonstructural components in general, there are restrictions on the maximum
value that can be selected for vy (or for the minimum value forK). From this discussion it
should be clear that lateral displacement (lateral drift) plays an important role in the EQRD
of structures, and that there is a need to establish the maximum value that can properly be
selected. This is not an easy task.

From the above discussion it becomes clear that to achieve economical EQRD the E. must
be minimized to the value required to achieve a serviceable structure. The problem, then
is to find out what can be done to minimize E. .

Combining and rearranging the above

equations results in the following expression for the energy stored in the structure.
Es

= E. - (E( + EJ

(1. 5)

Although, strictly speaking, for any given structure undergoing given ground motions the
input energy depends on the behavior (elastic or inelastic) of the structure, for the sake of
simplicity it may be assumed in the first approximation that
behavior.

E. is independent of structural

Then, considering that to achieve economical EQRD the energy stored in the

structure must be minimized, it is clear from Eq. 1.5 that it is necessary to dissipate a
significant part of the total input energy either by viscous damping, by inelastic
deformations, or by a combination of the two.

Although the advantages of controlling the seismic response of civil engineering structures
by increasing the equivalent viscous damping have long been recognized, the concept of
using plastic deformation of the structural material to dissipate part of the seismic input
energy does not appear in the technical literature of the United States until the 1950s. In
1956, Housner discussed the use of limit design for EQRD [1]. The use of the concept of
ductility and ductility ratio in EQRD of reinforced concrete (RC) structures was introduced
in the United States for the first time in 1961 with the publication of the Portland Cement
Association (PCA) Manual, "Design of Multistory Reinforced Concrete Buildings for

-3Earthquake

Motions" [2].

Since the publication

of the PCA Manual,

significant

experimental and analytical research efforts have been devoted to the development of
reliable methods of EQRD based on a combination of strength and ductility. In 1969, the
advantages of using plastic methods for the EQRD of ductile moment-resistant steel frames
was noted [3], and in 1975 and 1977 computer programs for the application of inelastic
design methods for the EQRD of ductile moment-resistant RC frames based on the use of
design ductility were developed and proposed for use in structural design practice [4 and 5].
Despite these accomplishments, the practical application of inelastic EQRD in the United
States is not widely used. This also seems to be the case worldwide, except for countries
such as Mexico and New Zealand, where building codes have explicitly introduced the use
of ductility ratio, J.L, in the estimation of seismic design forces, and allow the use of limit
design methods. In New Zealand, the seismic code is based on combined limit design and
capacity design procedures which incorporate the concepts of strength and ductilit).

The slow progress in the use of limit design and capacity design procedures for EQRD is
not surprising. The definition of the term "ductility ratio" and its evaluation are precise only
for the case of ideal linear elastic-perfectly plastic behavior, which, in reality, is the
exception rather than the rule. Furthermore, even though the advantages of providing a
structure with the largest ductility that is economically feasible are generally recognized, the
term "ductility" is used very loosely to express either the deformability of the structure or
the ductility ratio. Although the deformability, ductility and ductility ratio are interrelated
parameters, their values and significance in the actual behavior of structures can be quite
different. There is an urgent need to get a worldwide agreement regarding the proper use
of these technical terms and their proper evaluation and application to EQRD of structures.

1. 2 OBJECTIVES
The ultimate goals of this report are to review the state-of-the-practice and the state-of-theart in the use of the concepts of ductility, ductility ratio, and drift for attaining efficient

EQRD, and then to identify the research, development and educational needs required to
improve the proper use of these concepts.

-4-

1. 3 SCOPE

To achieve these goals, we will discuss both the need for applying the concepts of ductility
and control of drift and their proper use. Initial emphasis is placed on the importance of
recognizing the differences and interrelationships between deformability, ductility, and
ductility ratio, and on the need to unify the ways in which the different types of ductility
ratios are estimated from the actual seismic response of structures. It is then shown that
to achieve high energy dissipation capacity and overall effective seismic performance, it is
necessary to utilize highly redundant combined structural systems with multiple lines of
structural defense.

It is also necessary to provide the critical elements which control the

inelastic behavior of these systems with the highest ductility ratio that is economically
feasible.

The needs for controlling lateral deformation are discussed next.

Different

definitions of drift indices are presented, and their relationship to strength and ductility are
discussed.
After a brief statement of EQRD problems, the state-of-the-practice and the state-of-the-art
in the use of the concepts of ductility and drift limits are reviewed. The review of the stateof-the-practice focuses on how the concept of displacement ductility ratio (J..L,) and interstory
drift index (IDI) are used in present building seismic codes. Although emphasis is placed
on the Japanese and U.S. codes, the New Zealand, Mexican and European seismic codes
are also analyzed. Results obtained from the analyses and critical comparisons of these five
codes are summarized and discussed.

The state-of-the-art in the use of J..L, and IDI is

reviewed by assessing the implications of lessons learned from recent earthquakes and
research results. Comparisons of the state-of-the-art and the state-of-the-practice are used
to identify more fully research, development and educational needs. Short and long term
solutions are formulated for the proper use of the concepts of J..L, and IDI for EQRD of
buildings, placing emphasis on the need for an energy approach.

-52. NEEDS FOR DUCI1LITY AND CONTROL OF DRIFT AND THEIR


PROPER USE IN ESTABliSHING REUABLE EQRD CRITERIA

2. 1 NEEDS FOR DUCI1LITY

2. 1. 1 General Remarks It is well recognized and accepted that in EQRD all structural
members and their connections and supports, i.e.,all critical regions whose yielding strength
may be reached and exceeded by a severe earthquake, should be designed (sized and
detailed) with large ductility and stable hysteretic behavior so that the entire structure will
also be ductile and display stable hysteretic behavior. There are two main reasons for.this
ductility requirement:

first, it allows the structure as a whole, through distribution of

internal forces, to develop its maximum potential strength, which is given by the combination
of the maximum strength of all components; and second, large structural ductility allows the
structure to move as a mechanism under its maximum potential strength, which will result
in the dissipation of large amounts of energy. (It should be noted that attempts should be
mad, to develop a mechanism of a complete nature as opposed to a partial or local
mechanism, as in the case of soft stories.) While these two reasons have been recognized
in the past, only the second has been emphasized because the large dissipation of energy
was used to justify the reduction of the design strength that would be required if only linear
elastic behavior were permitted. Although this reduction is justifiable in certain cases, the
authors have previously expressed their concerns [6-9] about excessively large reductions in
the required elastic strength or in the Linear Elastic Design Response Spectra (LEDRS)
through the indiscriminate use of large values for the structural ductility ratio. For clarity
and convenience in discussing the reasons for this concern, a glossary of the terms to be
used in the discussion is given below.

Deformability: Capability of a material, structural component, or entire structure to deform


before rupture.

-6Ductility:

Capability of a material, structural component, or entire structure to undergo

deformation after its initial yield without any significant reduction in yield strength.

Ductility Ratio or Ductility Factor, 1-':

The ratio of the maximum deformation that a

structure or element can undergo without a significant loss of initial yielding resistance to
the initial yield deformation.

The above definitions are illustrated in Fig. 2.1 for the case of a Ductile Moment-Resistant
Space Frame (DMRSF).

2. 1. 1 (a) Needs for Recognizing the Differences Between Deformability. Ductility and
Ductility Ratio. Although the ductility ratio depends on the ductility, which in turn depends
on the plastic deformability, so that the three terms are interrelated, there are essential
differences in their quantification that need to be recognized.

Defonnability vs. Ductility: While one structure can have significantly greater deformability
than another, its ductility (particularly its usable ductility) can be smaller. For example, this
can be the case of a very flexible RC-DMRSF vs. a stiff but very ductile shear wall. It is
clear from analysis of Fig. 2.2 that if the DMRSF is too flexible, i.e. the
the maximum lateral deformation,

~f"''

.dfy

is very large, and

, that can be accepted or tolerated is limited, then

the DMRSF ductility that can be used could be smaller than the available and usable shear
wall ductility.

Ductility vs. Ductility Ratio: The difference between these two terms is clearly illustrated
in Fig. 2.2. While the shear wall usually has smaller ductility than a DMRSF, it can have
a significantly higher ductility ratio.

2. 1. 1 (b) Advantages of Providing Structural Components and Their Connections with the
Largest Ductility Economically Feasible.

The minimum desirable ductility for each

component should be such that the structure has the opportunity to develop its maximum

-7-

potential strength according to the maximum strength of its components. The need for this
is illustrated in Fig. 2.3, where the strengths of a simple structure composed of a ductile
moment-resisting frame and two coupled walls are depicted as the sum of the resistance
functions of each of their components. This figure illustrates that, in order for a structure
to develop its maximum potential strength RT as determined by the sum of the maximum
strength of each component (RT =

R.... + R,..2 + RF), it is necessary that

J..., 1 2.. 4.3, J. 912 2_2.8

and J.i.F 2_1.0. To allow the structure to move as a mechanism under its maximum potential
strength, the ductility ratio of the walls, particularly wall

w.,

must be significantly higher

than that of the DMRF. This figure also illustrates the difference between ductility ratio
and deformability. While the ductile moment-resisting frame has a larger deformability than
the walls, its ductility ratio can be smaller than that of the individual walls. The available
frame ductility ratio cannot be used effectively because the frame has significantly larger
deformability (flexibility) than the wall components do, resulting in a relatively earlier
failure of the wall components.

It should be noted that by providing large ductility and due to three-dimensional

(3-D)

interaction between DMRF and walls, it is possible that the maximum strength of the entire
structure will exceed the sum of its components if the strength of each is determined
considering it as acting independently.

This is illustrated in the schematic representation

(Fig. 2.4) of the behavior observed in the experiments conducted on the 7-story RC
DMRSF-wall structures of the U.S.-Japan Cooperative Research Program. Results of these
experiments are discussed in detail in Ref. 6. The beneficial 3-D interaction was identified
as a consequence of the effects of outriggering action of frames on the wall, as illustrated
in the isometric view of Fig. 2.5. The wall, rocking around the compressive edge during its
ductile axial-flexural behavior, tends to lift up the surrounding girders of the DMRSF that
frame into the walls. The girders resist this movement and, in doing so, develop higher
shear forces and increase the axial compression in the wall, which results in an increase in
its axial-flexural capacity. This outriggering action results in a significant enhancement of
the lateral strength of the whole structure.

-82. 1. 1 (c) Quantification of the Ductility Ratio. Though the use of the concept of ductility
ratio for the EQRD of structures was introduced in U .~.earthquake engineering literature
in the early 1950s, though its application to RC structures was presented in 1961 in the PCA
Manual [2], and though tremendous experimental and analytical research efforts have since
been devoted to its evaluation and application, even today it continues to be an ambiguous
parameter.

In a workshop conducted in 1977 [10], a group of experts, including professors,

researchers and practicing engineers, after recognizing the need to survey, analyze and
evaluate the main parameters

(as well as the definitions) that are presently used in

analytical and experimental research and in practice to describe the inelastic mechanical
characteristics of reinforced concrete materials, sections, regions, members, subassemblages,
structures and whole soil-building systems, made the following statement:

"One parameter of particular concern is ductility. While ductility is a useful


concept, it has a precise definition and quantitative meaning only for the
idealized case of monotonic, linear elasto-perfectly plastic behavior. Its use
in real cases where behavior significantly differs from this idealized case leads
to much ambiguity and confusion.

It is thus difficult to make valid

comparisons of "available" ductility values reported by different researchers


because they are often based on different response parameters or on yielding
values determined

using different and/or unexplained definitions.

These

experimentally obtained "available" ductility values are also often misused in


analytical studies of the "demand" or "required" ductility due to the difficulty
of establishing realistic values for the "linear-elastic stiffness and yielding
strength." Attempts should be made to integrate the definitions of response
parameters that are used in experimental test programs and in analytical
investigations.

Furthermore,

it is highly questionable

whether

the

performance of different building systems can be properly described and


evaluated on the sole basis of elastic stiffness, yielding strength, and ductility.
Consequently,

there is a need to introduce additional

parameters

for

describing the total hysteretic energy dissipation, number of cycles of reversed

-9-

deformations, and the degradation in stiffness and strength that has been
observed under seismic conditions."

The needs stated above are still valid today.

2. 1. 1 (d)

Concluding Remarks.

While in discussing the philosophy of ductility-based

design, it is possible to use the concept of ductility and/or ductility ratio in a vague manner.
But when such philosophy has to be applied in the EQRD of real structures, the philosophy
has to be quantified, and therefore it is necessary to use unambiguous parameters which can
be evaluated numerically with reliability. Such parameters are usually the displacement
ductility ratio, JJ., , and/or the rotation ductility ratio, J.l.e Preliminary designs are usually
based on a selected maximum JJ. 6 , which is determined based on the maximum values of JJ.e
which can be developed or which can be accepted at the critical regions of the structural
members.

Assuming that the values of JJ. 6 can be selected and evaluated reliably, the problem that
remains is to use this ratio or parameter correctly in the design process of a structure. To
discuss the solution of this problem it is advisable to review briefly the state-of-the-practice
and state-of-the-art in EQRD of RC structures. This is done after the discussion for the
needs to control IDI.

2. 2 NEEDS FOR CONTROLLING INTERSTORY DRIFT INDEX (IDI)

2. 2. 1 General Remarks In Ref. 11 it is stated that

"While displacement ductility factors generally provide a good indication of


structural damage, they do not usually adequately reflect the damage to
nonstructural elements.

This is an important limitation in seismic-resistant

design since a significant portion of the hazard to occupants and of the total
cost of repairing earthquake

damage is a consequence of nonstructural

-10damage.

Nonstructural

damage

is more dependent

on the relative

displacements (drift) than on the overall displacements. To obtain a reliable


measure of nonstructural damage, maximum drifts must remain unnormalized
or be divided by the value of drift corresponding to the damage threshold.
Nonstructural damage estimates based on drift ductilities may be misleading.
For example, nonstructural damage for relatively rigid structures may be small
even for large values of displacement ductility since the yield displacement
may be well below the nonstructural damage threshold. On the other hand,
the non structural damage and lateral displacements for flexible structures may
become intolerably large even before significant yielding develops.

To produce safe and economical structures, seismic-resistant design methods


must incorporate drift (damage) control in addition to lateral displacement
ductility as design constraints."

For clarity and convenience, in discussing the importance of control in the earthquakeresistant design of any engineering system, and particularly in case of building structures, the
following glossary of terms to be used in the discussion is introduced.
illustrated in Fig. 2.6 [12] and Fig. 2.7 [11].

2. 2. 2 Drift Definitions
Drift: Relative lateral displacement between two points (two floors).

Overall Drift = A .... (Fig. 2.6)

lnterstory Drift

= A ; - A ~~

Overall Drift Index lnterstory Drift Index -

11

(Fig. 2. 6)

""' (F.Ig .2. 6)

Ai-Ai-1

h .

(Fig. 2. 6)

The glossary is

-11-

Tangential-Interstory Drift: = (4 ; -4 ... )... = Drift-Producing Damage (Fig. 2. 7)

Tangential-Interstory Drift Index: (IDI)...

= RT (Fig. 2. 7)

2. 2. 3 The Need for Drift Design. As discussed in more detail in Ref. 12, the control of the
drift of a structural system under earthquake excitation is important for at least three
different reasons:

(1) to maintain architectural integrity, thereby avoiding unacceptable

damage to nonstructural components; (2) to limit structural damage and avoid structural
instability (P - 4) problems; and (3) to avoid human discomfort under frequent minor or
even occasional moderate earthquake shaking.

Story drifts and drift ductility factors may also be useful in providing information on the
distribution of structural damage. Unfortunately, conventionally computed story drifts may
not adequately reflect the potential structural or nonstructural damage to multistory
buildings. In some structures, a substantial portion of the horizontal displacements results
from axial deformations in the columns. Story drifts due to these deformations are not
usually a source of damage [Fig. 2.7 (a)].

A better index of both structural and nonstructural damage is the tangential story drift
index, RT . As schematically indicated in Fig. 2.7 (b), the intent of this index is to measure
the shearing distortion within a story. For the displacement components shown in Fig. 2. 7.
(c) and assuming that floor diaphragms are rigid in their own plane, the average tangential
drift index is equal to
RT

= -H(u-u )
-:Jl

L (u6 +

+ -

tL
Cl

u_
- u4''
-z

(2. 1)

in which L is the bay width and H is the story height. This first term on the right-hand side
of Eq. 2.1 is the conventional story drift index, and the second is a correction applied for
each bay accounting for the slope of the floors above and below the story. It may not be

-12appropriate to average the values of RT for a story when the pattern of axial column
deformations varies greatly across the structure (e.g.,frames with structural walls).

Although drift indices and, in particular, tangential drift indices, provide a good measure of
the distribution of structural deformations, it may be difficult to compute corresponding
yield drifts. One possible method for computing the yield drift is taking the drift present
at the appropriate location when the building, loaded with equivalent seismic lateral forces,
reaches its yield displacement; another is computing a story shear-drift relationship for a
subassemblage consisting of the story in question with appropriate boundary conditions.

The general philosophy of EQRD for structures (particularly for building structures) other
than essential facilities has been well established and proposes:

(1) to prevent nonstructural damage in frequent, minor earthquake ground shakings;

(2) to prevent structural damage and minimize nonstructural damage in occasional


moderate earthquake ground shakings;

(3) to avoid collapse or serious damage in rare major earthquake ground shakings
(where structure could be damaged but should not collapse).

The above philosophy is in complete accord with the concept of comprehensive design.
However, current design methodologies fall short of realizing the objectives of this general
philosophy [13]. A typical hierarchy of earthquake limit states is shown in Table 2.1 [14].

As pointed out by Dowrick [14], even up to the present it has been common practice to
design normal structures or equipment to meet only the two criteria designated above as (1)
and (3). Indeed, design usually has only been carried out explicitly for criterion (3), on the
assumption that the other two criteria (particularly criterion (1)) would be satisfied
automatically.

-13The growing concern over the costs of earthquake damages (direct, functional and indirect)
and the difficulty of repairing much postyield damage, points out the need for more
attention to be given to control of damage and repairability at the design stage. These
needs have been clearly emphasized by the 1989 Lorna Prieta earthquake.

The control of

damage will, of course, also help to improve human safety, which is the traditional
fundamental criterion. The main source of damage is deformation; thus to control damage,
it is necessary to control deformation and particularly to control interstory drift.

In summary, achievement of reliable and efficient EQRD requires satisfaction not only of

the criterion for strength and toughness but also the criteria for deformation

and

repairability. It should be noted that strength, toughness, deflection control and repairability
are interrelated and hard to define.

Assuming that the values of the IDI can be estimated reliably, the problem that remains is
how to use correctly this ratio or index in the design process of a structure. To discuss the
solution of this problem it is beneficial to review briefly the state-of-the-practice and stateof-the-art in earthquake-resistant

design of RC structures.

Tahle 2.1 IIIERARCIIY OF LIMIT STATE DESIGN CRITERIA FOR DIFFERENT LEVELS OF EARTHQUAKE HAZARD [14]
(A)
Serviceability
limit state

'Usual' design
cart hqua ke

Serviceability
eartiHJuake
(or OBE)
Response
condition

(I)

Undnmaged
Elnstic

Normal
st rue! ures
or
equipment

Critical
st rue! urcs
or
equipment

(2a)

No collapse
Post-yield cycling
Limited deformation
(Repairable)

(2b)
Typical
return period (yr)
Response
condition

Survivability
limit stale
Survivability
earthquake
(or SSE)
(3)

5-10
or TA"'" T 11 15 (?)

VI

50-100

500-1000
(5)

(4)

Pre-yield

500-1000

Pre-collapse

pre-yield

Typical
ret urn period (yr)

(C)

(D)

Ultimate
limit state

As (2a)
or
Pre-collapse

5000-10 000

-17-

R (RESISTANCE)

~F,

~w

~Wult

A (DEFORWA TION)

Fig. 2.1 - DEFINITIONS OF DEFORMABILI'IY,


DUCTILilY, AND DUCTILilY RATIO
R (RESISTANCE)

~ :~EAR

~ REAL 3-0
.EW.
BEHAVIOR

...a.:ALL
,.. - -

~.

..,

SUM OF INDIVIDUAL BEHAVIOR


OF WALL AND DMRSF

A (DEFORMATION)

Fig. 2.2 - DEFORMABILilY AND DUCTILI1Y


OF AN RC WALL AND AN RC DMRSF
R (RESISTANCE)
.._._ _ _ DEFORMABILITY OF DNRSF

----1

DUCTILITY RATIO (I'- ~1~.,) OF OMRSF

~F.

Fig. 2.3

A(OEFORMATION)

DUCTILI'IY REQUIREMENTS FOR BOTH WALLS AND FRAMES IN AN


RC FRAME-WALL SYSTEM

-18-

R (Resistance)

-----

w2 2 8

SHEAR WALL 1

P.w2 6 7

- - - - - . . , S H E A R WALL 2

p.F 1

p.F-2~

1-H--#-.,__-!--..l,---~--,._.....----------r

~w1 y 6w2 y

Fig. 2.4

D UCTJLE MOMENT
RESISTING FRAME
(DMRF)

f1 (Deformation)

THE EFFECTS OF A 3-D INTERACTION ON THE STRENGTH OF AN RC


FRAME-WALL SYSTEM

Fig. 2.5 - ISOMETRIC VIEW OF WALL ROTATION


ILLUSTRATING 3-D OUTRIGGERING EFFECT

-19-

r
II.

!"

-r

-rl
h.

!'

. Fig. 2.6 - DEFINITION OF DRIFfS (12)

60
H

,-DISPLACEMENTS
RESULTING FROM
'

CHORD

(A

ROTATION},~

f-:-L-tf

r--

i -

A )

i-1 T

r-----L----

DRIFT PRODUCING
DAMAGE

~ . ; ._~,.-~

HORIZONTAL ++------

STORY DRIFT

Cal DRIFTS DUE TO


AXIAL F~CES IN
FIRST STORY COLUMNS

(b) DRIFT DUE TO


STORY DEFORMATION

(c) DISPLACEMENT

COt.PONENTS FOR
COMPUTING R

Fig. 2.7 - COMPUTATION OF TANGENTIAL-INTERSTORY DRIFf INDEX [11]

-21-

3. STATE-OF-THE-PRACTICE

AND STATE-OF-THE-ART

OF EQRD OF RC STRUCfURES

3. 1

PROBLEMS

IN DESIGN

AND CONSTRUCTION

OF EQ-RESISTANT

STRUCfURES
The problem areas have been identified and discussed in detail by the authors in a series
of publications [7-9]. Only the three main problem areas that have been identified are
enumerated herein. The first problematical area in EQRD is in establishing the critical
earthquake
determining

input (Design Earthquakes).

The second includes problems involved in

the demands on the entire soil-foundation-building

nonstructural components) system by the critical earthquake.

(superstructure

and

The third involves the

visualization (for preliminary design) and prediction of the real supplies to the building at
the moment that an earthquake strikes.

The supplies and demands, in general,, involve the mechanical characteristics of stiffness,
strength, stability, and energy absorption and dissipation capacities.

Evaluation of the

demands and prediction of the supplies are not straightforward. The determination of the
demands, which is usually done by numerical analysis using mathematical models of the
entire soil-foundation-building system, depends on the interaction of this system as a whole
with the excitations that originate from changes in the system environment.
determination also depends on the intimate interrelation

betw~

The

the demand and supply

itself. Specific problems encountered in the three problematical areas of the earthquakeresistant design of structures -- critical earthquake input, demands on the building, and
supply capacities of the building --are discussed in Refs. 7-9.

While a sound preliminary structural design and reliable analyses of this design are
necessary, they do not ensure an efficient earthquake-resistant

structure.

The seismic

response of a structure depends on the state of the entire soil-foundation and superstructure
system at the time that earthquake shaking occurs, which means that the response depends
not only on construction, but on maintenance as well. A design will be effective only if the

-22model used can be constructed and maintained.

Although the importance of construction

and maintenance in the seismic performance of structures has been recognized, insufficient
effort has been made to improve these practices --through, for example, supervision and
inspection.

3. 2 STATE-OF-THE-PRACTICE
This review will focus only on the state-of-the-practice of EQRD of buildings as reflected
by present building seismic codes, and will emphasize how the concepts of J. 6 and IDI are
used and/or how they could be used to improve the state-of-the-practice according to
present knowledge.

Before discussing and comparing the seismic codes for the United

States, Japan, New Zealand, Mexico and Europe, it is convenient to make the following
general remarks regarding code estimation of demands and supplies.

3. 2. 1 Estimation of Demands in Present Seismic Codes. Although the review has focused
on U.S. seismic codes, the problems identified below are common to most codes in the
world.

There are several sources of uncertainty in code-specified procedures for the

estimation of demands, uncertainties that can be grouped into two categories: (1) specified
seismic forces; and (2) methods used to estimate response to these seismic forces.

3. 2. 1 (a) Strength For regular buildings, statically equivalent lateral seismic forces can be
derived as follows.

V = C.W =(C.., I R) W

For base shear:


where V is base shear,

c. is the

(3. 1)

design seismic coefficient, W is the weight of the reactive

mass (i.e., the mass that can induce inertial forces),

c.p is the

seismic coefficient equivalent

to a Smoothed Linear Elastic Design Response Spectral (SLEDRS) for acceleration, S., (C.p

= C.R = S/g),

and R is the reduction factor. Although in most of the codes the values of

R are given without any explicit relation to

J. 6 ,

these values implicitly depend on JJ. 6

-23-

Structural response is usually estimated using linear elastic analyses of the effects induced
directly by the above statically equivalent lateral forces or by these forces multiplied by load
factors depending on whether the design will be performed using allowable (service or
working) stress or the strength method.

There are only very few countries in which the

codes recommend or encourage the use of limit analysis and limit design methods.

3. 2. 1 (b) Stiffness and Drift. Most of the seismic codes address design for lateral stiffness
and for drift at service level. There are only a few codes that address estimations of the
change in stiffness during the response to major shaking and the maximum drift that can
occur.

There are also only a few codes that explicitly require that the contribution of

torsion should be considered in estimating the maximum lateral drift, and very few give any
guidelines regarding how to deal with the effect of multicomponent seismic excitations [15].

According to ATC 3-06, the deflections

o.

at ultimate (safety) limit state should be

evaluated as follows:
(3. 2)

where

cd

the deflection amplification factor (this factor is specified for different types
of structures)

oxe

the deflection determined

by an elastic analysis for the seismic forces

prescribed by the code considering the building fixed at the base

The ATC 3-06 [16] limits the maximum IDI, and consequently
hazard group to which the building belongs.

o., according

The limits vary from 0.010 to 0.015. For

certain types of buildings and limited height, a one-third increase in IDI


allowed if there are no brittle types of finishes.

to the seismic

limitations is

-24-

i
I

3. 2. 1 (c) P-.1 Effects. Few codes give explicit requirements and/or recommendations
regarding how to estimate these effects. While ATC 3-06 [16] recommends that the P-.1
effect does not need to be considered when the stability coefficient e is less than 0.10
(3. 3)

9
where
.1

= the

V,

= the seismic shear force acting between level x and x-1

h,.

= the story height below level x

P,

= the total unfactored vertical design load at and above level x

design-story drift

In accordance with this method the commentary to ATC 3-06 recommends that when

a is

greater than 0. 10, the design story drifts be multiplied by a factor 1I ( 1 - a) > 1. NEHRP-88
[17] requires multiplication by a factor 0.9 [11(1-a) > 1.

The 1988 UBC has a very vague requirement regarding P-.1 effects. This code requires
checking whether at working load level the ratio of secondary moment ( == Pw
primary moment ( == V w

c5 X&) to

h) does not exceed 0.1 (where Pw is the total unfactored [working]

dead and live load; V..., is the seismic shear at working load level; and h is the height of the
story). There is no doubt that there is a need for more rational code procedures to estimate
the demands regarding the stability effects at ultimate limit states.

3. 2. 2 Estimation of Suwlies in Present Seismic Codes.

3. 2. 2 (a) Strength. Most of the RC EQRD codes require that the supplied strength be
estimated

using a strength method in which nominal strength of critical sections are

I
\

-25-

evaluated as a function of just the minimum specified strength of the materials, and then
reduced by a strength reduction factor. There are a few codes, like the New Zealand Code,
in which the design and detailing of the critical regions of the structure is based on the
probable supplied strength capacity to the members.

Although most of the present RC

EQRD codes specify minimum size and reinforcement detailing according to the ductility
ratio that is expected to be developed, this is done in an implicit way. It can be concluded
that the state-of-the-practice, as reflected by most of the present EQRD codes for RC
buildings, does not appear to have included the use of the concept of energy dissipation
capacity in a rational and reliable way through the use of the ductility ratios.

3. 2. 2 (b) Stiffness. Deformation and Stability Capacities. Most of the RC codes give only
empirical expressions to estimate the so-called "effective linear elastic stiffness." Some of
them, such as the 1988 UBC which is based on the ACI (318-83) [19] (with some additional
provisions), include the effect of cracking in the estimation of the effective linear stiffness.
However, the UBC, as well as most of the present codes, does not specify how to evaluate
the changes in stiffness of the whole soil-foundation, superstructure and non-structural
components system induced by increasing damage.

There is a need to develop code

procedures that, based on the supplied local energy dissipation capacity of the structural
members (rotational ductility ratio and degradation with repeated cycles, i.e. local hysteretic
behavior), will lead to the estimation of the global deformational capacity of the structure
under not only monotonically increasing deformation but also under generalized (repeated
reversal) deformation.

The so-called Tri-Service Technical Manual [20], which will be discussed in more detail
later, recommends a method called "Method 2; Capacity Spectrum Method" which requires
estimating the "Force-Displacement Capacity Curve." Although the way that this curve is
used is questionable, there is no doubt that the requirement to estimate this curve is a step
forward toward solution of this issue.

-26-

3. 3 STATE-OF-TIIE-ART IN DUCfiLITY AND STORY DRIFT-BASED EQRD


The state-of-the-art with respect to each of the problem areas identified above is discussed
in detail in Refs. 8, 9, and 21. Only the state-of-the-knowledge regarding the proper use of
the concept of ductility ratios and story drift in the EQRD process will be discussed herein.

It is well recognized that EQRD requires an iterative procedure in which a preliminary

design is improved through a series of analyses. The importance of a proper preliminary


design should be highly emphasized, because, if the design procedure is started with a poor
preliminary design, the only thing that will be achieved at the end through its repeated
analyses will be an improved bad design. Before discussing separately the state-of-the-art
in

J.J- 6

and in IDI, some general remarks about their use are presented.

3. 3. 1 General Introductory Remarks Regarding Importance of u. and IDI in Preliminary


Design. Although it is generally well recognized that damage is due to deformation, there
is no agreement regarding what is the main criterion for preliminary EQRD of structures.
Perhaps as a consequence

of past and present code requirements,

present practice

emphasizes the use of strength in the preliminary design of structures. More specifically,
in most of the present codes, the preliminary design is based on only base shear strength
with a requirement to check the drift by elastic analysis. The insistence on only using
strength as primary criterion is perhaps a consequence of the following two reasons: first,
the practice of trying to design for ultimate or safety limit state by reducing the actual
inelastic design to one at working stress where linear elastic analysis can be used; and
second, the assumption that there is a unique relation between strength and stiffness without
recognizing that (particularly in the case of reinforced concrete structures) it is possible to
change significantly the strength of a structure without changing its stiffness.

While

preliminary design based on base shear strength could be justified for design where
serviceability (elastic response) controls, it cannot be accepted for cases where the design
is controlled by the ultimate (safety) limit state where large plastic deformation is accepted:
at this limit state, base shear strength of a given designed structure is insensitive to variation
of deformation and, therefore, to damage. Once structures yield, the base shear strength

-27-

remains constant, while the deformation can take any value, from its yielding value up to
the maximum value, at which collapse (sudden significant drop in resistance) occurs. In
view of the above insensitivity of the base shear strength to damage in the inelastic (plastic)
range of response (behavior), there have been some proposals to base preliminary design
on only lateral stiffness, i.e. on only controlling the interstory drift. However, a practical
method for this type of design has yet to be developed.

The authors believe that the most rational approach is one which recognizes the importance
of strength and stiffness (control of deformation) and which also recognizes that these two
factors, while strongly interrelated in the case of elastic response, clearly have a weaker
relation to each other in the case of inelastic response.

In this last case, the lateral

deformation at ultimate limit states depends on the yielding strength provided to the
structure. To control inelastic deformation, it is necessary to provide the structure with a
minimum yielding strength. Therefore, to achieve an efficient preliminary EQRD, there is
a need to consider two requirements simultaneously the strength (based on the rational use
of

J.1. 6 )

and the deformation (based on the limitation of IDI).

3. 3. 1 (a)

Examples.

The following two simple cases emphasize the need to consider

strength and deformation simultaneously.

Given (See Fig. 3.1):


(a) structure: SDOFS [Fig. 3.1.(a)]
(b) ground motion: an impulsive EQ-ground motion, consisting of either
a rectangular acceleration pulse [Fig. 3.1. (b)] or a
triangular acceleration pulse [Fig. 3.1. (c)], the two
which can be considered respectively as extreme

cases

regarding probable shapes of pulse.


(c) intensity of pulse: service level O.lOg = (vJ.
Maximum Credible EQ

= MCEQ, 0.40g

of

(vJ.

-28Required:

To design a structure whose maximum deformation

will not exceed the

following acceptable deformations.

(v)ocrvia:

0.005h(v) .....

0.02h

Solution
FIRST CASE: Rectangular Acceleration Pulse [Fig. 3.1. (a)]

Assume a td I T: to be specific, say td I T = 514

Design for Serviceability: Linear elastic response. In this case, for td I T ;:: 0.5, the value
of the maximum Dynamic Load Factor (DLF) ..... is 2.0. See Fig. 3.2 [22].

Strength (Resistance) Required:

(R,.j..,,_.;"" = 2

(O.lOg) m

(R,.joervioe = 0.20 W

Lateral Stiffness Required:

Knowing the mass m, and value of the selected period T, it is

possible to compute the required stiffness k from

T = 21t

lm ~ k = (21ti m

~k-

(3. 4)

T2

Knowing k, it is possible to compute the service displacement [(v..rvi"")


to compare with the required limit for (IDI).

= (v..rvioe)lh,

say (IDI).

= R_,.lk]
0.005

and then

-29-

> (voervia:)e>:>ep~ablc

If (v oervia:)c~ooman:~

stiffness has to be increased. This will induce a change in T

(decrease) and a new iteration in the design procedure has to be started. Assuming that the
preliminary design for serviceability was satisfactory, then we can go to:

Design for Safety: [(vJ...J = 0.40g. Assuming that the supplied Ry


linear elastic plastic response and for

= 80 =

for a

I T = 514 from the graph (Fig. 3.3) with a

R,/F 1

the graph indicates a v"'.)v.1..,;.

= R, = 0.20W,

J.L 6

= 0.20W/(0.40g)m = 0.5
This is a ductility ratio which not only cannot be

developed but which also results in a vmax = 80 vclastic = 80 0.005h = 0.4h, i.e. an (IDI)~ X

0.4, which obviously cannot be accepted.

From the above computations, it is clear that there is a need to:

(i) increase the yielding strength; or

(ii) increase stiffness; or


(iii) a combination of (i) and (ii).

Increase in stiffness will lead to a decrease in the (v...J..rv1"'' but as the T will decrease the
td I Twill increase and then the v"'/v.lasti will increase. Thus the solution is to try to increase
the strength; but by how much? It depends on how much p, 6 can be developed economically
and on how large (IDI)max can be tolerated to control damage.
Assume that l.h < 6 and (IDI) < 0.02. From Fig. 3.3 it can be seen that to have a p, 6 =
vmax/vclastic = 6

-+

R,/FI

= 1.05,

then R, = 1.05 0.40g m = 0.42W. Note the significant


X

increase in strength that is needed (0.4210.20) = 2.1.

Check lateral stiffness. If we keep the same k as for the service design,
vmax = (vclast;.)max

6 = 2.1(vclastic), 6, i.e. (v...J = 12.6(vclasti.)..rvicc


X

-30If the (vewuc)oervia: i.e. under the F 1effcclive

= 0.20W, was already

0.005h, the (v.....J

= 0.063h. This

results in an IDI which is too high. Usually the maximum acceptable is 0.02. Thus to limit
the IDI it will be necessary either to decrease acceptable J. 6 or to increase stiffness or a
combination of these two solutions.
For example:

doubling k (which will reduce the period T to a value equal to Tl ./2 and

therefore
td
--=1.4 td {f=l.8;

(3. 5)

TJ/2
and reducing J. 6 to 4 will result in a required yielding strength of
R,..

= 1.15, 0.40W = 0.46W.


vmax

Vmax

Then

::;: ( 0.46)( Velastic acceptable]


2
0.20

::;: 4 .6(0.00Sh)
(3. 6)

0.023h

The resulting IDI of 0.023 may be acceptable.

SECOND CASE: Triangular Acceleration Pulse [Fig. 3.1 (c)]

It can be legitimately claimed that the occurrence of a rectangular pulse is practically

impossible.

To demonstrate that the same type of design is required even by the less

demanding type of acceleration pulse, which can be considered to be the triangular pulse,
the above example is repeated for this kind of pulse and considering a
td I T = 1 rather than 1.25, and assuming T = I sec.

Design for Serviceability

For td I T = 1 and (vJ..,"';"" = O.lOg, from the graph of (Fig. 3.2), the

-31{R,.jelastic oervioe = 1.5(vJ..rvi.,.m = 0.15W


and
v..rvi.,.= {R,.,J.~asuc ..rvi.,./k = 0.15W/(2rrYm/(1 ,y = 0.15g/39.48

Assume h = (294) in. Then IDI = 1.47 in/294 in

= 1.47 in

=0.005.

Design for Safety

F.trec~ive = 0.40W

Assuming R = R, = (R,.Jscl'\;a: from the graph (Fig. 3.4) it can be seen that for
td I T = 1, and

R,,./F.rrecti-.: = R,.,.IF. = 0.15/0.40 = 0.375,

the required J.La = va.)v.lasti<

= 22

This value is too large to be developed, and also results in an IDI = 22


nearly five times more than can be accepted.

0.005 = 0.11,

Then, as discussed previously in the

improvement to the solution of the FIRST CASE, the supplied yielding resistance has to be
increased, either by a decrease of the acceptable J.La. and/or by an increase of the lateral
stiffness.

Assume that k is doubled; Tis decreased by T/ v'2 and consequently ~ I T = 1.41. With
this value of td I T and considering a J. 6 = 6, the required R,,.,/F.trective
0.28W and

vmax = (vewtJ0.28W

6 = (0.005/2) h

(0.28/0.15)

6 = 0.028

which, although closer to 0.02, still is not acceptable.

= 0. 7.

Then R_. =

-323. 3. 2 (b) Concluding Remarks Although none of the above assumed ground motions are
realistic representations of usual EQ shaking, the examples illustrate very clearly the need
to consider simultaneously the strength (through the proper use of JJ. 6) as well as the
deformations (through the acceptable limits of IDI). Similar procedures can be applied in
the case of more realistic EQ ground motions. The only basic information that is needed
is the:

SLEDRS for service EQs,


SIDRS for maximum creditable EQ,
Acceptable limits for IDI at service
as well as at safety limit states.

where SIDRS

3. 4

= Smoothed Inelastic Design Response Spectra

STATE-OF-TilE-ART

IN USING DUCTILITY RATIO JJ., IN PRELIMINARY

EQRD

According to the previous discussion, J1. 6 should be used throughout the whole design
procedure, and particularly in the final step, where the final designing and detailing of the
critical regions of the structures are done. However, as the importance of this last step
concerns other parts (phases) of the whole research project, only the use of JJ., in estimating
the demands will be discussed herein, specifically in: (1) establishing the design EQs, and;
(2) the preliminary design of the structure.

3. 4. 1 Use of u. in Establishing the Design EOs.

The following two main different

methods are being used.

A.

REDUCTION

OF THE LINEAR ELASTIC DESIGN RESPONSE SPECTRA

(LEDRS) THROUGH THE DIRECT USE OF THE VALUE OF JJ. 6 (NEWMARK


AND HALL PROCEDURE [23]), OR THROUGH THE USE OF STRUCTURAL
MODIFICATION FACTORS, R, (ATC-3 PROCEDURE).

R is a function not only

-33of J.J., but also of the provided overstrength, OVS, and of the increase in damping, ~.
due to large deformations.

B.

DERIVATION

OF IDRS

THROUGH

STATISTICAL

STUDIES

OF THE

INELASTIC RESPONSE SPECTRA (IRS) OF STRUCTURES TO AVAILABLE


RECORDED OR EXPECTED (PREDICTED) CRITICAL GROUND MOTIONS.
These IRS are obtained through time history nonlinear dynamic analysis of structures
with different yielding strengths (Cy , or different degrees of J.J.,) and

o [24].

This

method can be considered as a part of the overall energy approach to the EQRD [25,

26].

Method A, which is very simple, is already widely used and has been included in codes of
several countries. However, as the proponents of this method point out, the method is only
valid for very limited types of structures. The application of this method to the design of
most real buildings is questionable [24-27].

Method B can be considered as the method of the future.

Although it has already been

applied to simple cases, its general application in practice will require extensive integrated
analytical

and

nonstructural

experimental

studies on real

component systems.

3-D soil-foundation-superstructure

and

Once a reliable IDRS has been attained, the next

problem is how to use J.J., in the preliminary design of the structure.

3. 4. 2 Use of u. in Preliminary Design. For the purposes of this discussion, the different
ways of conducting the preliminary design of EQ-resistant structures as regards the use of
the ductility concept in the sizing of the structural members can be classified in the following
three groups:

(a) #1a is not used at all. The critical internal forces in the members are

obtained through linear elastic distribution (LED) of forces; (b) implicit and partial use of
#1a

Usually this is done by allowing a limited amount of redistribution of the internal

moments that have been obtained through a LED of forces; (c) use of limit design
approach.

Different methods, varying from the one based on simple plastic theory, which

-34assumes infinite ductility, to those based on a more general plastic theory which consider
"linear elastic serviceability conditions," as well as realistic limitation of IJ.e and
incorporating

also stability considerations.

J, 6 ,

These methods are usually classified as

compatibility and serviceability methods, with serviceability methods being the most
promising of the two. This group also covers methods that include the possible occurrence
of shakedown phenomena, which are, at present, being developed.

3. 4. 3 Concluding Remarks. In summary, the authors believe that the future of EQRD is
an energy approach in which the concept of ductility is used by combining the methods B
and c, i.e., Be, with proper consideration of the possibility of shakedown phenomena.
However, this is considered a long-term approach. In present practice, as will be discussed
later, most of the methods that are used can be classified as under the combination Aa.
Although methods that can be classified as combined Ab are being used and have been
investigated recently [28-30], the results of these investigations indicate the need for further
studies regarding:

(1) the proper limits in the amount of redistribution; and (2) the

adequate redistribution pattern through the height of the structures.

In view of the above remarks, and the fact that it is usually not practical to change radically
the state-of-the-practice, the authors would like to formulate for the immediate or very near
future the following compromise solution: to conduct the preliminary design using improved
Ab or Ac (or even Aa) methods; but ideally this should be complemented with time history
nonlinear dynamic analyses of the response of the preliminary designed structure to the
predicted maximum credible EQ (MCEQ) ground shakings that can occur at the site of the
structure during its service life. If this is not possible, the least that should be conducted
is a static nonlinear analysis of the response of the structure under a monotonically
increasing lateral load. Before this compromise solution can be applied in practice, it is first
necessary to identify the improvements needed, and then to carry out the studies required
to achieve these improvements.

-35-

3.5 STATE-OF-TIIE-ART IN USING INTERSTORY DRIFT INDEX


(IDI) IN PRELIMINARY EQRD
It is perhaps unfortunate that in the past most of the efforts in improving EQRD has been

expended only in designing for strength without proper consideration


deformation.

Damage is a consequence of deformations.

of the role of

For any structure that is

responding in the inelastic (plastic) range under practically a constant strength, the degree
or level of damage depends upon the amount of the plastic deformation that the structure
undergoes. Thus, to control damage it is necessary to control deformations. The question
is how to achieve such control at the different levels of EQ shaking that can occur during
the life of the structure.

Although the general EQRD philosophy contemplates at least

three different limit states, for a preliminary design it usually is convenient to try to satisfy
only the first (serviceability) and the last (safety) limit states.

3. 5. 1 Control of IDI at Serviceability Level. Because at service level the structure should
respond in its elastic range, then for a structure with a given damping and a total mass and
its distribution, the control of IDI depends on the stiffness. The question is what is or are
proper limits for IDI to control nonstructural damage as well as to protect other contents
of the building and perhaps avoid human discomfort. At present it is generally recognized
that the limits for IDI depend on the function (occupancy) of the building, type of
nonstructural , elements and how these nonstructural

components are attached to the

structure. Depending on these factors, acceptable limits for IDI vary from 0.0006 to 0.006.
Although the estimation of the IDI at the service level is usually based on linear elastic
analyses, there are many uncertainties regarding the effective stiffness of the structural
members (particularly on RC structures due to cracking), the deformation of the foundation,
the effects of the nonstructural components, etc. Furthermore, analyses of the deformation
should be based on a realistic 3-D model that considers properly the effects of torsion under
the multicomponents of ground motions. The estimation of torsional deformations offers
serious difficulties.

-36The most practical procedure for including the criteria of IDI control in the preliminary
design is to use the SLEDRS (tripartite on logarithmic scale or even the standard SLEDRS
for displacement, Sd on normal scale). If the limit IDI is specified assuming uniform IDI
along the structure (or, according to type of structure, considering possible concentrated IDI
at top or bottom story by multiplying average IDI by a factor smaller than one), the
maximum acceptable displacement can be estimated and from the SLEDRS the maximum
acceptable period of the structure can be found. This preliminary estimation of the period
facilitates the selection of the PSa that has to be used for the design for strength. Once the
preliminary design for strength has been completed, a realistic 3-D model of this preliminary
design can be used to estimate the maximum IDI that can occur under the multidirectional
EQ ground motion.

3. 5. 2 Control of IDI at the IDtimate or Safety Limit State. The acceptable maximum IDI
to control damage varies with the type of structure and the function of the structure.
Usually it varies from 0.01 to 0.03. The IDI demands for different types of EQs can be
estimated by evaluating the IDRS for different values of J.L, or Cy. The problem is how to
estimate the effective lateral stiffness or period T.

3. 5. 3 Choice of Member Stiffness for Drift and

P-~

Analyses. [12] A common difficulty

in seismic analysis of RC structures is the selection of a set of rational stiffness values to be


used in force and displacement analyses. Should one use gross concrete-section properties?
Should one use some reduced section properties?

Or should the gross concrete properties

be used for one type of analysis and reduced section properties for another?
design codes in the U.S. are not specific about this matter.

The seismic

Thus, the choice of section

properties used in lateral analysis in general, and in seismic analysis in particular, varies
widely.

Contributing to the complexity of this issue are the following factors:

-371.

Although elastic material behavior is usually assumed for the sake of simplicity,

reinforced concrete is not a homogeneous, linearly elastic material.


2. The stiffness and idealized elastic material properties of a reinforced concrete section
vary with the state of behavior of this section (e.g. uncracked, cracked, yielding, and ultimate
states).

3. Not all RC members in a structure, and not all cross sections along a particular member,
are in the same state of behavior at the same time.

4. For many beams and other asymmetrically reinforced members, the stiffness properties
for positive bending and negative bending are different.
..
"'

5. Uncertainties regarding the actual contribution of the floor slab and of the deformations
of the joints.
6. The stiffness of reinforced concrete members and structures varies with the time, and
with the history of past exposure to wind forces and EQ ground motions.

7. The stiffness of RC members and structures varies with the level of deformations.

Analytical and experimental studies reported in Ref. 31 indicate that for motions which are
within the working-stress design limits of members, the measured fundamental periods of
concrete structures are generally slightly less than the periods computed using grossconcrete-section properties.

According to these studies, in the case of large-amplitude

motions up to the yield level, the stiffness of the building is usually somewhere between the
computed values based on the gross concrete-section properties and the cracked-section
properties. Based on this observation, the same reference suggests that for force analysis,
the gross concrete-section properties and the clear-span dimensions be used, and the effect
of nonseismic structural and nonstructural elements be considered. For drift calculations,

-38either the lateral displacements determined using the above assumptions should be doubled,
or the center-to-center dimensions along with the average of the gross section and the
cracked-section properties, or one-half of the gross section properties should be used.
Furthermore, the nonseismic structural and nonstructural elements should be neglected if
they do not create a potential torsional reaction.

Similar sets of assumptions have been proposed by research workers who have been
concerned about the choice of member stiffnesses to be used in the P-A analysis of concrete
structures.

For example, for second-order analysis of concrete structures subjected to

combinations of gravity and wind loads, MacGregor and Rage (32) recommend using 40%
of the gross section moment of inertia for beams and 80% of the gross section moment of
inertia for columns.

3. 5. 4 Recommended Practical Methods for Designing Considering IDI. A simplified


method to estimate lateral drift of RC structures has been suggested by Sozen in Ref. 33.
The method is intended to be used for interpreting experience and evaluating relative merits
of different structural schemes and member sizes on the basis of tolerable damage criterion.
The method is conveniently. used in preliminary evaluation by simple estimates of the base
shear capacity coefficient.

In Ref. 34 and 35 the effects of strength and stiffness and of the type of ground motion on
nonlinear displacement response of SDOF systems is investigated.

The results obtained

show that nonlinear displacement response is equal to the linear response spectral values
if the system has certain strength which is determined by dimensionless parameters for
strength, initial period, and type of ground motion.

Recent results obtained by some of the authors (36) have' shown that the nonlinear
displacements are very sensitive to the dynamic characteristics of the ground motions and
in some cases the displacement can be significantly higher than those computed from a
linear elastic response, particularly if high JJ. 6 are used in the derivation of the yielding

-39strength (see Fig. 3.5 for the case of Corralitos and Hollister ground motions recorded
during the 1989 Lorna Prieta EQ). This observation agrees with the results obtained by
Kappos in Ref. 37. These observations also agree with results reported in Ref. 38, which,
based on results obtained, recommend the use of the following empirical formula for
estimating the deflection amplification factor Cd (defined as the ratio of absolute maximum
interstory displacement to the corresponding value from a linear time history analysis).
In

cd

= 0.414 (1, J.L.J

(3. 7)

where J.l.m is the maximum story ductility.

In Ref. 39, Hatamoto et al. propose a newly automated seismic design method for RC
frames which aims at a uniform energy dissipation throughout the building frame, so that
the resulting damage is uniformly distributed as much as possible over all elements.

3. 5. 5 Need to Consider the Effect of Multicomponent Seismic Excitation and Direction


in Estimating Structural Response @D. At present most seismic code procedures as well
as traditional dynamic analyses of 3-D buildings assume that the applied excitations are in
the direction of the structure's reference axes.

For a real complex 3-D structure, it is

difficult to find the structural principal axes, and the centers of structural mass, stiffness and
resistance may not be coincident.

Thus, the direction of structural reference axes chosen

may not coincide with the structural principal axes. Therefore, when applied excitations are
in the direction of structural reference axes, the structural response may not be maximum.
This has been clearly illustrated and discussed in Ref. 15. In this reference it is shown that
the seismic components and their input direction can significantly affect the response of a
torsionally sensitive structural system.

Ground components applied at the structural

reference axes may remarkably underestimate the response, because the structural maximum
response is dependent on the seismic input direction and its magnitude.

._ - - m
h

\i gmax

vglllliX

____.
... t

l :__--~----~

- - - - ---

'\

'

J.. - - - . .

(c) TRIANGULAR
ACCELERATION PULSE

(b) RECJ'ANGULAit
ACCELERATION PULSE

(n) SIJOFS

t .. 7'\

Fig. 3.1 SINGLE DEGREE OF FREEDOM SYSTEM


SUUJECfED TO GROUND-ACCELERATION PULSES

I.6

2.0

I
1.6

I ,~i
'V /?
~ ~
1/
y

......J

1/v

:---

0.4

v
I--

0
0.05

VI
I
0.10

r.b_

0.5

'r~lr

1.0

0.2

10

r--....

"" ~

' '" '"

0.4

1"-

'/

"lli_
I
I

0
2.0

f-

;J

0.6

'L

0.8

lrt

I
I

....J,..
1\

lrt

0.2

......,o

-0.8

1.0

---

I. 2

v
b

1.2

"\

I
1/

I. 4

1.0

Fig. 3.2 LINEAR ELASTIC RESPONSE SPECTRA [22]

3.0

4.0

-42-

'
' / / / /Y/

7z;r/f

".:I

WI

I I I
I

0. I
0.1

I I I

1..0

10

40

Fig. 3.3 RESPONSE SPECTRA FOR Al~ ELASTO-PLASTIC


SDOF SYSTEM: DUE TO RECTANGULAR PULSE [22]

-43-

I
~0

; .

I ;

I 'II I

/,

I I VI I I

I :

20

'

I I /1

. \/1 I I 1/

1
I

: 1/ ., 1 I 111
lA Ill I I/ l

I I II 17 :7

. 'I
I
I
r-:-- .f-ll_ ~ L -L _y_

I I

17'

.l /J I l I 1/
l.E Tl. _/..{ 1f- +V-4 -tT _jI VI' / . . . v I
II
I

!
- -10.8 _l_

.__H itfTitftV:~J-r-Jtr+-

~~~~/~-~~-~~~~~~~--~~
1
8 ~~~~~+1-4~~~~1
I
1 I
I
I /
I /1 I 1 /
I
I
I
I
I I I I I
I
I I . I I I y I ;r I (A
I~ I
I Tl I I I I I
I

5 ~~~~~~~~+-Y*V~+.Y_I~Y~T~~~~~~~~~~~~~41~1~14-~~~
!/1 ~I II VI/ I~ I I ; II I I
I 09 ~

1 l!i v

v 1

lf-~/l_j -Y~/-~y
/:

1.0

t '

1 1 1

/1

r
I

/-;7

Y Y A

::

I / 1....-(' I I I I I

/ ' / Y L/1 /1

I I I I II

~;~y:~'Yl\

v~r--~

0.2~-l-LL

'-~"':

~ '-i
1 ""-.1

I~ ~:

'Y'Ii;

1.20-+--

A I "!/I~
?1'""'-1 _...;......_ -:;( I

1.60~

,.!___

p' r- ~tziQO:
~ u_
F
~ !.(
)',{
L1

t.J

(111

P.esislonce
function

0.1

vr ,

;m~~~~,t~~:~llf u.~ai~

0.8 //>

o.s

/n ~

7 r4' L...-1- ~ L J _JJL1_Ll_-i _l


~ I "'
I .--I 011 II
' l

f-

Oisclacement
function

I
l
irianqulor culse
l
1.._~1~~~-l..'----l.'__,.!-,1~~p_odl....,l.!... I'--_:_'-l-7-'--'-..!...'.-..:..'~...J......J:......;'W'....!',..--~1---:i

0.1

0.2

0.5

!-J'

0.8 1

10

20

'

Fig. 3.4 - RESPONSE SPECTRA FOR AN ELASTO-PLASTIC SDOF


SYSTEM DUE TO A TRIANGULAR LOAD PULSE [22]

-44-

CORRALITOS

DISP.(in]

90 DEG

~----------------------------~~~

14

~=5%

--j.J.=l
--- J1. = 2
---- J1. = 4
----- J1. = 6

12
10

--------

8
6

2
0

0.5

0.0

DISP.(inJ
14
12
10

2.5

2.0
1.5
1.0
PERIOD (sec)

HOLLISTER

--J.l.=l
--- J1. = 2
-- - - J1. = 4
----- J1. = 6

3. 0

DEG

90

~=5%/
I

,,

.,.

,..-- .

'
'---' /

"'

8
6

2
0
0. 0

0.5

1.5
2. 0
1.0
PERIOD (sec)

2.5

3. 0

Fig. 3.5 NONLINEAR DISPLACEMENT RESPONSE SPECTRA [36]

-454. STATE-OF-TilE-PRACTICE:

REVIEW OF CODES

4.1 COMPARISON OF TilE BASIC SLEDRS: UNITED STATES' UBC AND JAPAN'S
BUilDING STANDARD LAW (BSL)

Comparison of the Basic SLEDRS adopted by UBC and BSL for the static lateral load
procedures, which are illustrated in Fig. 4.1, shows that:

4. 1. 1 Shape of the SLEDRS The two above codes assume that, for low values ofT, there
is a constant value for the amplified acceleration (or response pseudo-acceleration)

which

neglects the decrease in the amplification of the acceleration to one as the period of the
structure tends to zero. While UBC adopts a sharp transition from the constant value of
the pseudo-acceleration to the decreasing values, the BSL adopts a smooth decreasing curve
which seems a more rational transition.

4. 1. 2 SLEDRS Values Comparing the values for the three standard type of soils (soil
profile types) S1 , S2 , and S3 , it is clear that except for low values of the Fundamental
Period (T

Natural

0.6 sec.) and values ofT > 2 sees. ,the spectral values specified by BSL are

higher than those required by UBC. The differences in the case of soil profile type S3 (soft
soil) reach significant values. For example, in case ofT = 1.6 sec., the value of the BSL is
1.28/1.6

= 0.80. This is 1.46 times

the value specified by UBC, which is 0. 75/(1.6)213

= 0.55.

Only the special soil profile S4 introduced by UBC compares well with the values of the S3
given by the BSL (see Fig. 4.1).

4. 2 COMPARISON OF TilE DESIGN SPECTRA

4. 2. 1 Japan's BSL. This is the only code of those analyzed herein that explicitly specifies
two-level design earthquakes:

moderate earthquake ground motions, which may occur

several times during the use of the buildings and should be resisted with almost no damage;
and severe earthquake ground motions, with a low probability of occurrence during the use
of the buildings and which should not cause collapse or harm human lives.

-464. 2. 1 (a) Service-Level Design Earthquake. The stresses caused by moderate earthquake
ground motions shall not exceed the allowable stresses for temporary loads.

These

allowable stresses are: for concrete, two-thirds of the specified concrete strength; for steel,
the yield stress. The lateral seismic shear coefficient Q; of the ith story above ground level
shall be determined in accordance with the following formula.

(4. 1)

where C; and

w; are

the lateral seismic shear coefficient of the ith story as determined in

accordance with the formula in Eq. 4.3, and the weight of the reactive mass above the ith
story.

For base shear at service level:

(Qs).

= (V s),

= (Cs).(WJ,

(4. 2)

(WJ. =The weight of the building, which shall be the sum of dead load and the applicable

portion of live load, and

(C.),

= Z.R,.A; Co = Z.R.(1)(0.2)

(4. 3)

where
Z =

the seismic hazard zoning coefficient, which for Zone A (the zone of highest seismic
risk in Japan) is

1.0.

R = the design spectral coefficient (see Fig. 4.1) which is similar to the C factor = C of
UBC.

A;

the lateral shear distribution factor (equal one at the base).

Co

the standard shear coefficient assumed to be 0.2 for moderate earthquake motions.
This defines the SLEDRS at service level.

-47Based on above Eqs. 4.1 to 4.3, the base shear for the zone of highest seismic risk is:

(Qs).

= (Vs). = (C.),(WJ. = R.(0.2)(WJ.

(4. 4)

(C.). = 0, 2 R. defines the SLEDRS at the service limit state and is shown in Fig. 4.2.

4. 2. 1 (b) Ultimate or Safety Level Design Earthguake.

Japan's BSL specifies that limit

analysis design procedures be used to ensure the minimum resistance against severe
earthquake motions.

For this level earthquake, the ultimate lateral shear strength of each story shall not be less
than the necessary ultimate lateral shear Qun =; V.., determined in accordance with the
following formula:

Vun

=Qun

=D. F.,. QLd

(4. 5)

where: W R is the weight of the reactive mass under severe ground motions, and C" is given
by Eq. 4.3,arid for severe earthquake motions (i.e. at the ultimate or safety limit state) the

Co is 1.0. Then C" is equal to

Cu

= {J.O){R..)u = (C.)u

(4. 6)

i.e., the direct spectral values given the SLEDRS shown in Fig. 4.1.

D.

the structural coefficient that reduces (de-amplifies) the SLEDRS (represented by


R. in Fig. 4.1) due to the dissipation of energy that takes place due to inelastic

behavior which depends on the available ductility. For RC with excellent ductility
0.3

D,

0.4. If it is assumed that D. = 1/J.L 6 , then 3.33 > J.L 6

2.50.

-48F" =

Fe. F. where: Fe is a Shape Factor that depends on

= the eccentricity of stiffness

of each story (Fe varies from 1 to 1.5); and F.= is a Shape Factor that depends on
the variation of R. = the lateral stiffness, and varies from 1 to 1.5.
For the sake of simplicity, let us assume that Fe = 1. and Fce = 1., then

Vun = Q.,

= (C.)u WR = (0.3)(RJuWR

1 for RC-SMRSF

(4. 7)

(4. 8)

The corresponding smoothed design spectra for the safety level of these two types of RC
structural systems are shown in Fig. 4.3.

4. 2. 1 (c) Comparison of Japan's BSL and UBC Design Earthquakes:

Service limit State

Level. Because UBC does not specify design earthquake at service level, it is not possible
to compare the two codes.
IDtimate or Safety Limit State Level. The design seismic coefficient (C.)u at this level
specified by BSL can be obtained from Eqs. 4. 7 and 4.8 and is graphically shown in Fig. 4.3.
According to BSL this Smoothed Inelastic Design Response Spectra (SIDRS) should be used
with limit analysis to design structures. The term "limit analysis" is used herein to denote
an analysis in which is allowed redistribution of internal forces due to plastic actions.

For structures other than for essential facilities the UBC permits damage that will not
jeopardize human life. Neither UBC nor BSL defines explicitly what constitutes acceptable
damage.

UBC recognizes that the level of acceptable damage has to be different for

different types of facilities depending on the type of occupancy (Occupancy Categories).


This is done quantitatively through the use of the Importance Factor, I. However, as it will
be discussed later, the values of I, which vary only from 1 to 1.25,do not appear to be able
to cover the entire range of damage that can be tolerated for the existing different types of

-49occupancy.

For example, let us consider Structures to Shelter Essential Facilities or

Hazardous Structures (Categories I and II, respectively). These structures should survive
severe (extreme) earthquake ground motions without any significant nonstructural and
structural damage, i.e., without yielding; thus, their response should remain in the linear
elastic range. If this is so, the seismic forces for which they should be designed against are
those specified by the SLEDRS.

However, according to UBC, the structures for these

essential and hazardous facilities can be designed for seismic forces that are significantly
smaller than those required by the SLEDRS. Specifically for the case of RC-SMRSF, the
UBC allows them to be designed for

Seismic Forces = 1.25(SLEDRS/12) U

i.e.

(0.104)(SLEDRS)(l.4) = (0.146)(SLEDRS)

= (SLEDRS)/7

(4. 9)

In other words, it is allowed to de-amplify the SLEDRS by a factor of 117. (As it was noted
previously, the de-amplification adopted by BSL is only 1/3.3.) Although the authors agreed
in principle with the following explanation given in the Commentary published by SEAOC
[40],

"In the judgment of SEAOC the details of design and construction often
dominate seismic performance.

Therefore increasing these aspects for

essential facilities can improve performance more effectively than relying


solely on increased design force levels to achieve this end."

it is believed that to control damage it is necessary to design {provide) structures with a


minimum yielding strength and stiffness which cannot be provided by allowing the above
large de-amplification of the SLEDRS.

-50Japan's BSL, on the other hand, appears to accept the same level of damage independent
of the type of occupancy or function of the facility for which the structure is designed. The
reasons for this philosophy are not clear and require investigation.

UBC permits significant de-amplification of the established SLEDRS through the use of the
so-called "system quality factor," R.,.

"

_1 (Z/(1.25)S) W

Rw

rm

_1 (2.75Z[) W

(4. 10)

Rw

or
(4. 11)

As the reduction or de-amplification factor

reduces the seismic design forces to a

working stress level rather than to a first yielding strength level, to obtain the base shear at
the yielding level, it is necessary to multiply it by the load factor U, which for RC is 1.4.
Then
(4. 12)
It should be noted that this V" is used with the ACI (Ref. 318-89) strength method which

is not a capacity (limit analysis) method. However if it is assumed that the ACI strength
method is used optimizing the design in such a way that a sufficient number of critical
regions reach their plastic moment to form (develop) a collapse mechanism, it is possible
to compare directly the design base shear of Eq. 4.12 with that used by BSL, i.e. Eqs. 4.7
and 4.8.

As for RC-SMRSF and Shear Wall structural systems, the R,... values are 12 and 8
respectively, and therefore

-51-

V,

= ~ (SLEDRS) W = 0.117 (SLEDRS) W for RC-SMRSF

(4. 13)

12

and
4
V, = 1. (SLEDRS) W = 0.175 (SLEDRS) W for RC-Shear Walls
8

(4. 14)

The two equations are plotted in Fig. 4.3, which permits direct comparison of the UBC and
BSL design earthquakes for severe (extreme) earthquake ground motions. From analysis
of the curves shown in Fig. 4.3, it is clear that the seismic coefficients specified by the BSL
are significantly higher than those specified by UBC. For example, for Soil Profile S 1 and
a structure with a T = 1 sec. and having a RC-SMRSF structural system, UBC allows to
design for a strength of Vu = 0.059 W, while the BSL requires Vu = 0.192 W, i.e. near three
times that required by UBC. Furthermore, the Japanese require that W include some live
load, and this is not generally required by UBC. Note that if the ACI strength method is
used without any plastic redistribution and without attempting to design all critical regions
for just their required strength, the resulting designed structure can have significant story
yielding overstrength.

It appears that either Japan's BSL is too conservative in its seismic coefficient or the UBC

is not conservative enough. Note that even the seismic coefficients specified by the BSL for
moderate (service) ground motion (Fig. 4.2), are larger than the seismic coefficients specified
by UBC for severe earthquake ground motion. It appears that UBC uses

R...

values which

are too large for reducing the SLEDRS, and therefore implies the use of an unrealistically
high ductility ratio in reducing the required strength.

It should be clearly noted that it is not possible to judge the effectiveness of the earthquake-

resistant construction of any country just by looking at the code requirements of the seismic
forces to be used in the design. The quality and effectiveness of any earthquake-resistant
construction depends on the design philosophy (selection ofbuilding configuration, structure,

-52layout, and structural system) among many other factors or aspects, including the following
main ones:

1.

Code requirements which include not only those regarding the seismic forces, but even
more importantly, the minimum requirements for the materials, sizing, and detailing
of the structurai and nonstructural components, as well as inspection during all the
phases of design and construction.

2.

Construction aspects (construction technologies) - quality control of materials and


workmanship as well as field inspection.

3.

Maintenance (preservation).

The degree of damage during an earthquake depends on the deformation (stiffness) of the
structure.

To control damage it is necessary to control the deformation.

However, as

previously pointed out, the resulting deformations are related to the yielding strength
provided to the structure.

There is a minimum strength that has to be supplied.

If this

minimum strength is not available the inelastic deformations can reach unacceptable values.

4. 3 OBSERVATIONS REGARDING DERIVATION OF SIDRS DIRECrLY FROM


SLEDRS
The rationale of deriving SIDRS directly from SLEDRS through the use of ductility ratio
(J..L) values has been questioned [8, 9, 21]. It is recognized that the effects of dissipation of

energy through inelastic deformation can be considered as equivalent to the effects of


increasing the damping (i.e. the yielding damps out the engineering resonance phenomenon).
However, the effect of dissipation of energy through plastic deformation is not the same as
just increasing the value of the damping constant c or of the damping ratio

for a linear

viscous damping mechanism, which is usually considered in setting up the structural equation
of motion and solving it to obtain through linear elastic analysis the LERS for SDOF
systems. This is clearly illustrated in Fig. 4.4, where, for a given specific recorded ground

-53-

motion, the LERS of SDOFS under various damping ratios

(0% through 60%) is

compared with the 5% damped IRS obtained considering the effect of the dissipation of
energy induced by the inelastic (plastic) behavior of the SDOF when its yielding strength
(or its yielding seismic coefficient Cy) is reduced considering different values of the ductility
ratio J.L.
The above observation that "the reduction in structural response due to viscous damping
mechanism, i.e., due to an increase of damping ratio

~'is

not the same as the reduction

induced by early yielding due to increase in acceptable ductility ratio p." is also clearly
illustrated by comparing the spectra of these reductions as recommended

by several

researchers, illustrated in the graphs of Fig. 4.5 [41]. From these graphs it is clear that the
reduction due to J.L decreases as the frequency increases. On the other hand, the reduction
due to increase in

increases as the frequency increases (see Figs. 4.4 and 4.5, aas well as

Table 4.1, where it has to be noted that the displacement region of the spectra corresponds
to the lowest frequency and the acceleration region to the highest frequency). Therefore
it appears that any attempt to reduce the SLEDRS due to inelastic behavior (i.e. due to J.L)
by selecting a new SLEDRS corresponding to an increased value of

is questionable. This

is the main reason for questioning the procedure recommended in what is usually called the
Tri-Services Manual approach [20, 42].
4. 3. 1 Tri-Services ManuaJ Approach. This manual recommends that essential buildings
(as well as all other buildings) be designed to resist two levels of earthquake ground
motions. This is a step forward in improving design approach. The first level of motion is
designated as EQ-1 and the second and larger amplitude of motion is designated as EQ-11.
The lateral force-resisting structural systems of these facilities will be designed to resist EQ-I
by elastic behavior (i.e. using SLEDRS).

Then the facilities will be evaluated for their

ability to resist EQ-11 by post-elastic behavior with ductility ratio limitations.


procedures (methods) are recommended:

Two

-54-

4. 3. 1 (a) METHOD 1: ELASTIC ANALYSIS PROCEDURE.

This procedure evaluates

overstresses of individual elements, calculating the Inelastic Demand Ratios (IDR) which
are equal to the ratio of Demand Forces {M 0 , V0 , F 0 ) to the calculated yield or plastic
Capacities of the Elements <Me, Ve Fe). If the IDR exceed acceptable values, or some
other four conditions, the structure must be analyzed in accordance with Method 2. The
acceptable values for the IDR vary from 1 to 2.0 for Essential facilities; 1.25 to 2.5 for High
Risk facilities; and from 1.25 to 3.0 for all other facilities. These values seem reasonable,
however in the practical application of this method they are usually increased up to 10.

4. 3. 1 (b) METHOD 2: CAPACITY SPECTRUM METHOD.

A step-by-step approach

is recommended for use in estimating or approximating the inelastic capacity spectrum of


any given structure. Then this capacity is compared to the demands of the EQ-II response
spectrum.

Although the numerical procedure of estimating the capacity spectrum can be

questioned (particularly the determination of the change in the value of the fundamental
period or the effective period of vibration), there is no doubt that conceptually this
recommended procedure, which requires the estimation of the capacity of the structure to
resist lateral forces, is a step forward towards the improvement of earthquake-resistant
design methodology.

This method has gained rapid popularity among the professional

designers, as well as in some research [43].

Two seriously questionable procedures in Method 2 are: first, the recommended procedure
to estimate the demand curve which is obtained by estimating a reduction (or deamplification)

in response amplification

due to inelastic action; and secondly, the

comparison of the capacity spectrum with the demand curve. The reduction in response is
done by reducing the spectral values of the SLEDRS corresponding to EQ-II through the
use of a larger value of the viscous damping ratio

as it is illustrated in Fig. 4. 6. The real

problem with this method is that conceptually the actual SIDRS cannot be obtained from
a SLEDRS by only increasing the value

of~

up to 10% or even up to 20% as was discussed

previously and illustrated in Fig. 4.4. Usually this method can be quite conservative,
particularly for structures having a certain period T (T close to the value of the predominant

-55period [frequency] of the ground motion) and possessing a ductility displacement ratio equal
to or larger than 1.5. It is recommended

that to improve the present procedure, the

computed capacity response spectrum be compared with the demand spectra obtained by
estimating the IDRS corresponding to the Cy, as well as the
estimated lateral resistance-deformation

p.~

that results from the

capacity curve illustrated in Fig. 4. 7.

The direct comparison of the capacity spectrum (estimated according to the Tri-Service
manual) with the earthquake demand curve can be questioned because the estimated change
in the effective period of vibration can only become effective in reducing the response of
the structure in the event that the ground motion time history is of such nature that it
induces on the structure a response that consists of repeated cycles of deformation reverSals
whose peak intensity (magnitude) increases with time from the first yielding to a reversal
. requiring the maximum available displacement ductility [see Fig. 4.8 (a)]. The comparison
will be completely incorrect in the case of impulsive types of EQ ground motion, which
would require the structure to develop its deformation in just one cycle of inelastic behavior,
as is illustrated in Fig. 4.8 (b).

4. 3. 2

Comparative Designs of Buildings Using U.S. and Japanese Codes.

Several

comparative designs of RC and steel buildings have been conducted in cooperative studies
among Japanese and U.S. designers [44-49]. Because the Japanese BSL design procedure
is based on a specified story yielding strength, while the U.S. UBC design procedure is based
on the yielding strength of the member, it should be expected that the UBC would result
in design with higher overstrength. However, it is believed that this higher overstrength can
not overcome the significantly higher lateral strength required by the BSL. From the results
of these comparative designs it becomes clear that the current Japanese code requires
lateral strengths significantly higher (2 to 3 times) than the U.S. code. On the other hand,
in most of these studies it has been concluded that the drift requirements of the U.S. code
are more severe than the Japanese Code. This last conclusion needs further studies because
it appears that this conclusion has been derived by comparing only the drifts obtained under

-56-

different lateral load conditions and not under time history analysis of the response to the
same or similar ground motions.

4. 4

NEW ZEALAND CODE OF PRACI1CE FOR GENERAL STRUCfURAL

DESIGN AND DESIGN LOADING FOR BUJLDINGS (NEW ZEALAND STANDARD

NZS 4203: 1984)

4. 4. 1 Material Code Fonnat.

"Strength" design or "Limit States" design is used to

proportion members for design loading U. Then "Capacity" design is used to ensure that
during earthquake loading the preferred mechanism of energy dissipation is maintained.
That is, the action arising from "overstrength" under flexural or combined flexural axial and
shear stresses at the critical inelastic regions (usually called plastic hinges) are used as
design actions (forces) for other elements in the structure to avoid brittle failure elsewhere.

4. 4. 2 Material Strength. The "clepawbbJe strength" of members is calculated using the


specified (minimum) steel yield strength
strength reduction factor

fr

and concrete compressive strength f; and the

~.

The "CM:I::!I1tu:ug1h" used in capacity design is calculated using steel strengths of 1.25

fr

for

Grade 275 (Grade 40) reinforcing steel, 1.4 fr for grade 380 (55 ksi) reinforcing steel and
~=1.

4. 4. 3 Strength Reduction Factor. For strength design t = 0.9 to 0. 7 depending on action


(flexure, shear, etc.). For capacity design t = I.
4. 4. 4 Load Factor U. The required ultimate load capacity U is:

+ 1. 7 4
I. OOD + 1. 3 4 + E
0.90D + E

U = 1.40D

U =
U =

(4. 15)

-57-

where 4 =reduced service load.

4. 4. 5 Special Observations or As.pects. The capacity design procedure as used in New


Zealand is for buildings designed for ductile flexural yielding or for yielding of diagonal
braces.

4. 4. 5 (a) Earthquake Provisions. The NZS starts by stating a series of requirements that

are to be adopted to comply with the standard such as Symmetry and Ductility [which means
the ability of the building or member to undergo at least repeated and reversing inelastic
deflections beyond the point of first yield while maintaining a substantial proportion (;:::
85%) of its initial maximum load carrying capacity].

4. 4. 5 (b) Method of Analysis. According to the NZS, buildings shall be analyzed by the

equivalent static force method. However under certain special circumstances an additional
spectral modal analysis shall be conducted.

4.4.6 Equivalent Static Force Analysis: Total Horizontal Seismic Force or Shear: V, at the

base in each direction, shall be

(4. 16)

W,=D+C,L

(4. 17)

where: W, = total reduced gravity load above the level of imposed lateral ground restraint
and is given by the expression 4. 17, where C, varies from 0 to 2/3 depending upon the
intensity of the live load L; D = the dead load; and the seismic design coefficient

Cd = CSMR

(4. 18)

-58C

basic seismic coefficient, to be determined from Fig. 4.9in accordance with the seismic
zone (A, B or C), the subsoil flexibility and the period T.

Although not clearly

specified in the NZS, it appears that the spectral values for C can be interpreted to
represent the Design Earthquake (SIDRS) for the severe (extreme) earthquake ground
motions.

S -

structural type factor, intended to reflect the potential seismic performance of different
structural systems. The specified level of S (which varies form 0.8 to 5 for Reinforced
Concrete Structures

an~

0.8 to 6 for steel structures) primarily takes into account the

ability of the structural type to dissipate energy in a number of load cycles (i.e. sho.uld
reflect the available ductility), and secondarily its degree of redundancy where
appropriate.

M or Mp =

the structural material factor is a factor that allows consideration of the

effect of different materials in the potential seismic performance of different structural


systems. For example, for ductile prestressed concrete structures, M = 1.0 while for
ductile reinforced concrete, M = 0.8. This difference reflects the fact that in an
elastically responding structure, the one that is prestressed has a larger response (less
damping).
R

factor related to the relative direct risk to life of a failure [corresponds to the
importance factor (occupancy type or function of structure) I of UBC]. However, the
value of R can be as high as 2 while the largest value of I is 1.25.

Assuming R = 1, i.e., buildings with normal occupancy or usage, the seismic design
coefficient corresponding to a RC frame responding elastically and located on rigid or
intermediate subsoils will be given by the spectra shown in Fig. 4.10, i.e.

-59Ve

= Ce W, = (CSMR)

W,

= [C(5x0.8)

(1)] W,

= [4C]

W,

(4. 19)

For a similar structure but taking advantage of its maximum available ductility (actually
using its maximum capacity of energy dissipation) the design can be conducted using the
SIDRS Cu
V.,

= C.,W, = [Cx0.8x0.8]

W,

= [0.64C]

W,

(4. 20)

In other words the SIDRS C" can be considered to be obtained from the SLEDRS Ce by
reducing it by 4/0.64 = 1/0.16 = 6.25. Note that the NZS, similar to the UBC, reduces the
SLEDRS Ce by a constant value which is independent ofT. Furthermore, the NZS specifies
that for structures responding elastically, the seismic forces depend on the SLEDRS Ce, i.e.
V = 4C W, . However, for the design of structures containing highly hazardous content the
required V = (1.28C)W, , i.e., is obtained by reducing the SLEDRS Ce by 3.125. The
rationale for this reduction is not very clear and it appears to be incompatible with the fact
that these structures should remain practically elastic (i.e. respond in their elastic range).
Note that in the case of essential facilities (hospitals, etc.) the reduction is even higher
(3.91).

Similar incompatibility exists in the UBC. This code for hazardous and essential facilities
using RC-SMRSF allows reducing the SLEDRS by a factor of 12/(1.25 1.4) = 6.86. The
rationale for this large reduction needs to be investigated. While it is true that any small
amount of yielding (ductility) can reduce significantly the ordinates

of the LERS

corresponding to structures with period T equal or close to the predominant period of the
ground shaking T, , in general it is difficult to justify a constant reduction of the order of
3.91 in the case of the NZS and particularly of 6.86 in the case of UBC.

4.5 COMPARISON OF REQUIRED DESIGN SEISMIC FORCES AT SAFETY LEVEL


BY NZS WITH UBC AND JAPAN'S BSL
Figure 4.11 permits comparison of the required seismic forces at the safety (ultimate) level.
It should be noted that while NZS and the BSL permit the use of limit analyses, for RC

-60structures the UBC requires the use of the strength method which basically is not a capacity
method.

In general, for identical seismic forces, the application of limit design (based on

limit analysis) will result in a structure with smaller ultimate strength (or "overstrength")
than if it is designed just by the application of the ACI strength method. Only in the case
of "optimal strength design" (where every critical section is provided with just the demanded
strength) will this result in a structure with the same ultimate strength as that obtained from
applying limit design method.

It should be noted that the NZS limits the amount of

redistribution of the design moments obtained from an elastic analysis to 70% of the
moment for the section obtained from an elastic envelope covering all appropriate
combinations of loads; i.e. moment reduction should not exceed 30%. The ACI strength
method limits the redistribution (increase or decrease) by 20%.

Neglecting the effect of the amount of redistribution in the design method as well as
differences in the material sizing and detailing requirements among the three codes and
comparing only the required seismic forces as illustrated in Fig. 4.11, it is clear that:

1.

To account for the effects of inelastic behavior the three codes reduce the seismic
forces required for structures responding elastically through a reduction factor that is
independent of the period of the structure.

2.

The BSL requires the use of the largest lateral seismic forces; for structures with
T < 2 sec. the NZS requires the smallest forces. For structures with T close to 1 sec.,
the BSL requires forces which are more than 3 times the forces required by UBC.

3.

Of all the codes reviewed herein, the largest reduction of the seismic forces that are
required to obtain the design forces for safety to keep the response of the structures
linear elastic (i.e. to the SLEDRS) is permitted by the UBC (12/1.4 = 8.57) This
reduction is attributed primarily to the effects of dissipation of energy (ductility). The
NZS allows a reduction of 5/0.8 = 6.25 and the BSL is the one that specifies the
smallest reduction (110.3 = 3.3). In chapter 6 a more detailed discussion will be

-61-

offered of other effects, such as overstrength, which must be considered in order to


make a fair comparison among the design forces specified by the different codes.

It should be noted that seismicity m the highest seismic region of New Zealand is

significantly lower than that in the corresponding highest seismic region in the United States
and Japan.

This fact is clearly reflected in the SLEDRS adopted by New Zealand when

compared with that of the United States and Japan (Fig. 4.11).

4. 6 COMPARISON OF REQUIRED

LIMITS FOR THE INTER-STORY DRIFT BY

NZS, UBC AND BSL


The minimum required idealized performances of RC-SMRSF structures by the three codes
are presented in Fig. 4.12 for the case of a structure with a T = 0.31 sees.

4. 6. 1 NZS Requirements.

The inter-story deflections computed under the lateral seismic

forces specified at the safety level (multiplied by the factor K/SM, K being a factor by which
the values of C are scaled: K = 2.0 for the defined static forces, and 2.2 in the case of
dynamic modal forces) shall not exceed 0.0006 of the story height where non-structural
elements are not properly separated nor 0.010 of the story height multiplied by the zone
factor which varies from 1. 0 for Zone A to 2/3 for Zone C. The NZS does not have a
specific requirement limiting the drift under service seismic forces. However there is a
general building deformation requirement limiting the interstory drift index (IDI) to 0.002.
Using this value, the total lateral load (normalized with respect to the weight of the reactive
mass) interstory drift index relationship is shown in Fig. 4.12. It can be seen that to reach
the maximum IDI of 0.010 the structure will have to develop a ductility ratio of 5 which is
the value that the NZS appears to have adopted as feasible to achieve. However, it should
be noted that from the values specified for structural type factor S, it appears that the NZS
considers a maximum displacement ductility ratio of 5/0.8 = 6.25.

4. 6. 1 (a)

Calculations of Deformations.

It appears that the deflections are estimated

through a linear elastic analysis using the specified forces multiplied by K/SM (which for

-62RC-SMRSF is 3.125). If this is so, and considering that the maximum deflection at service
load is 0.002 h, it would appear that the designed structures can never reach the maximum

= 0.00625 h

0.01 h because (0.002 h)(3.125)

< 0.01 h.

4. 6. 1 (b) Building Separation. The NZS specifies that parts of a building, or buildings on
the same site separated from each other shall have a minimum clear space from each other
either 1.5 times the sum of their computed deflections or of 0.004 h, whichever is the larger
and in any case not less than 25 mm (1.0 in.).

4. 6. 2 UBC Requirements.

Calculated story drift shall not exceed 0.04/R,., nor 0.005 times

the story height for buildings with h

> 65 feet.

For buildings with h

> 65 feet, the

calculated story drift shall not exceed 0.03/R,., nor 0.004 times the story height.

The

specified story drifts are those due to the specified seismic design forces using linear elastic
analysis. Thus they are at the working stress level. The UBC does not specify directly the
maximum acceptable drift under the extreme earthquake ground motions.

However,

according to the UBC requirement for deformation compatibility, it is assumed that the
maximum displacement should not exceed 3(R,/8) times the displacement resulting from
the required lateral forces.. Thus the maximum interstory drift for SMRSF structures of
building with height :::; 65 feet should not exceed:

m~ (O~~l

= O.O!Sh

(4. 21)

and for R,., = 12

(i)

Rw (0.005h)

= 0.0225h

(4. 22)

This means that 0.015h controls.

Note that, according to the above requirement, if the structure is designed just to satisfy the
requirement that the IDI under the working lateral forces is equal to 0.04/R,., , then the

-63maximum displacement ductility ratio that can be used is 0.015/[(0.04/R...) 1.4] = 3.21, which
is smaller than the specified R... modified

to yielding level R,

= R,)1.4 = 8.57.

For

buildings with a height exceeding 65 feet, the serviceability requirements are controlled by
0.03/R...

= 0.0025, thus

deflection at yielding will be 0.0025 x 1.4

= 0.0035.

The maximum

acceptable deformation would be (0.03/R..,) x (3/8R...) = 0.0125. In this case the maximum
ductility that can be used is 0.0125/0.0035 = 3.57. Considering a building with T 2_1.64
sees. the minimum required strength would be: at working 0.03W; and at yielding 0.042W.
Then the idealized load-deformation relationship is that illustrated in Fig. 4.12.

4. 6. 3 Japan's BSL Requirements.

Under the service (moderate) earthquake ground

motions (i.e. V = 0.2R.W = 0,2 w for R. = 1.0) the story drift shall not exceed 0.005 of the
story height. This value can be increased to 0.00833 if the nonstructural members will have
no severe damage at this increased story drift limitation. Regarding the maximum interstory
drift under severe earthquake ground motions requiring a lateral shear strength of 0.30W
(see Figs. 4.3 and 4.11), to the best of the author's knowledge the BSL does not have any
direct or explicit specifications.
requirements

Considering that the deflections have to satisfy the

under the moderate earthquake ground motions and extrapolating these

requirements to the severe motions level we have Fig. 4.12. It is clear from this figure that
if the maximum interstory drift index should not exceed the value of 1.5% (as recommended
by ATC and UBC), the maximum ductility implied in the BSL design would be 0.015/0.0075

= 2 for

cases where nonstructural elements can be damaged, and 0.015/0.0125

1.2 for

cases where the nonstructural elements cannot be damaged. From these results it is very
clear that if Japanese designers try to satisfy only the minimum elastic stiffness required by
the BSL, they cannot take advantage of the maximum displacement ductility ratio that the
BSL appears to consider for reduction in SLEDRS for strength, Cr (i.e. 110.3 = 3.3).

It should be noted that from inspection of RC ductile moment frames constructed in Japan

it would appear that they are stiffer than required by the BSL. Furthermore, the above BSL
specifications apply only for buildings up to 60 meters (about 20 stories) in height. For
buildings higher than 60 meters (197 ft) special permission from the Ministry of

-64-

Construction following a detailed review of the dynamic behavior of the structure by the
Board of Technical Members is required. While it is true that the Japanese designers use
somewhat shallower girders than those used in the United States (to decrease story height),
the width of the floor girders is significantly wider than those used in the United States.
Furthermore, in Japan the span of the bays is usually smaller than in the United States.

4. 7 MEXICO:

1987 TECHNICAL NORMS (STANDARDS) FOR THE FEDERAL

DISTRICT
These standards recognize that acceptable damage is different for essential facilities (group
A) than for standard occupancy facilities (group B). However this differentiation is based
on adopting for group Aa seismic coefficient 50% higher than that for group B, or in other
words, it appears that it is not required to keep the essential facilities in the elastic range,
since this will require for group Aa seismic coefficient significantly higher than just 1.50
times the one adopted for group B. To keep the behavior of a RC-SMRSF for group A in
its elastic range requires a design with a seismic coefficient of about 4 times (value of Q')
the values adopted for the inelastic design of the same type of structure for group B (Fig.
4.13). Even in the case that for an essential facility it would be possible to accept structural
damage equivalent to a displacement ductility 2 (e = 2), which would demand considerable
damage at the most critical region of the structure, the seismic coefficient should be not just
50% but 100% higher than that for group B.

It should be noted that the Mexican Federal District Standards are the only ones of those
considered herein which specify in their static method of analysis that when the T of the
building tends to zero, the required strength seismic coefficient

c. tends

to the value of the

normalized peak ground acceleration (see Fig. 4.13).

4. 7. 1 Mexico Reguired Strength at illtimate (Safety) Limit State. The required maximum
strength design spectra for RC-SMRSF located in Zones I and III are shown in Fig. 4.13.
It is obtained from the SLEDRS by reducing it by a reduction factor of Q, which, contrary
to other codes, is period dependent.

While forT < 0.2 sec. in the case of Zone I (and T

-65-

< 0.6 in the case of Zone II) the SLEDRS is divided by Q' = 1 + (T/T.)(Q-,1), for any
value of T > T. = 0.2 sec. for Zone I and T > 0.6 sec; for Zone III the value of Q' = Q

=4

for the case of SMRSF. The reinforced concrete code requires that the design of

SMRSF be conducted using a strength method with a factored seismic load of 1.1 E. The
required strengths are shown in Fig. 4.13 for Zone III.

It should be noted again that from analysis of the IRS for different values of JJ. 6 , it becomes

clear that in general the required yielding seismic coefficient is not equal to the C required
by the SLEDRS divided by JJ. 6 The reduction varies with the T. This is clearly shown from
analysis of the values of the reduction for the Oakland Outer Harbor Channel 3 (Fig. 4.14);
while for value ofT > 0.5 sec. the reduction due to a JJ. 6 = 2 is higher than 2 (for T

=1

sec. the reduction for JJ. 6 = 2 is higher than 3). For T < 0.25 sec. and forT > 2 sees. the
reduction is smaller than 2. Furthermore, for values of JJ. 6 > 2 the reduction does not
increase in direct proportion to the values of JJ. 6 For example, for T = 2.0 sees. the
reductions for JJ. 6 = 3 through JJ. 6 = 6 are practically the same.

4. 7. 2 Mexico Requirements for lnterstory Drift. The 1987 Mexican regulations reqmre
that the maximum lateral IDI not exceed 0.006, except where elements that did not form
part of the structure were attached in such a manner that they would not be damaged by
structural deformations.

In this case, the limit of 0.006 is increased to 0.012. It should be

noted that these values include the inelastic deformation that was usually based on a Q
JJ. 6

= 4,

thus it appears that implicitly the Mexican code limits the linear elastic IDI to

0.0015 or 0.003, depending on whether the nonstructural elements can be damaged or not
due to the lateral deformations.

Fig. 4.12 illustrates the diagrams of expected lateral

strength-interstory drift index, according to the 1987 Mexican Code. From Fig, 4.12, it can
be seen that in the elastic range for either short or long T, the IDI requirements of the 1987
Mexican Code are the less severe except for the Japanese BSL. At the ultimate Limit State
of Safety and for short period structures, except for the New Zealand Standards, the limit
for the IDI is the most severe of all the regulations considered herein.

-66The 1987 Code also requires that all types of construction shall be separated from the
neighboring pieces of land at a distance not less than 5 em (2 in) or the horizontal (lateral)
displacement

calculated

for the floor level under consideration.

The horizontal

displacement should be computed with the reduced seismic forces (i.e. reduced by Q

= ~J. 6)

and then multiplied by the "factor of seismic behavior" Q = #- 6 , and the result increased by
0.001, 0.003 and 0.006 of the floor level above the ground for Zones I, II and III,
respectively.

4. 8 DUCTILITY AND DRIFT CONSIDERATIONS

IN EUROPEAN SEISMIC

CODES

4. 8. 1 Reference Documents. The following review is based on the following publications:


European

Recommendations

for Steel Structures

in Seismic Zones, ECCS,

Technical Committee 1, Working Group 1.3, Seismic Design, 1st Ed., 1988
Seismic Design of Concrete Structures, CEB, 1987.

Both documents advocate a very similar approach, but with a few important differences.
These differences are identified here wherever appropriate.

The ECCS notation is used in

this summary.

4. 8. 2 Summary of Relevant Aspects of Design Procedure:


4. 8. 2 (a) Seismic Input. Two primary quantities for defining the seismic input are:
Peak ground acceleration, A
Shape of elastic response spectrum, given by dynamic amplification factor for 5%
damping (see Fig. 4.15).

Recommended values for corner periods, plateau values, and shape of descending branch
vary between documents.

EECS recommends that the descending branch be proportional

to 1/T (constant velocity).

This branch is translated for three different soil types in a

manner similar to ATC 3-06 [16] for soil types S 1 ,S 2 , and S3

-67-

The CEB alludes to the use of different and site specific spectra, but the consensus in both
documents is that the ground acceleration A and elastic response spectrum shape are
adequate to describe the seismic demand for design purposes.

Duration effects are not

considered.

The ECCS recommends translation of the elastic response spectrum shape for damping
different from 5%, and also gives values for absolute and relative ground displacements to
be used in design.

4. 8. 2 (b) Design Spectrum. As illustrated in Fig. 4.16, a design spectrum shape is created
from the ground motion spectrum in a manner similar to ATC 3-06 [16], i.e.,the plateau is
extended toT = 0 and the liT line is changed to 1/(T)k, with a recommended value of 2/3
for k. The corner period

is not changed in the process (different from ATC). A lower

limit of 0.3 times the plateau value is recommended in the ECCS.

R(1)=Ro
Ro
R(1)=(I_)"

(4. 23)

To
R(1)~0.3Ro

This spectrum shape, R(T), multiplied by the ground acceleration A, is used as the basis for
deriving the seismic design forces. Again, the approach is similar to the R-factor approach
in ATC 3-06, except that the R-factors are called "behavior factors" and are denoted as q
(in ECCS) or K (in CEB). Thus the ordinates of the design spectrum C(T) are given as

-68C(T) A

R(T)
K

(4. 24)

In equivalent static force analysis, the C(T) values are used as base shear coefficients and
are distributed over the height in a manner similar to U.S. codes. In modal analysis, the
C(T) values are used as spectral values for SRSS superposition
In the design process, strength design is used with rather complex combinations of different
load effects. The seismic load effects show up usually without load factors, i.e. ,again similar
to ATC. A significant difference from ATC is that portions of the live loads are considered
to be seismically effective.

4. 8. 2 (c) Behavior Factors. These factors differ significantly between steel and RIC, and
are quite different from the R-factors used in the U.S.

For RC structures with ductility level III (highest ductility level), the CEB factors are

RC frame systems

K = 5

RC wall and dual systems

K = 4

For steel structures, the ECCS recommends the q-factors given in Table 4. 2. These qfactors contain the ratio

ex~/ cx 1

which is the ratio of lateral strength of the structure

associated with mechanism motion to lateral length associated with first yielding in the
structure. Thus the ECCS accounts for "mudme overstrength" due to redistribution in the
inelastic range.

4. 8. 2 (d) Drift Considerations.

Drift limitations are included in the documents primarily

for damage control, but reference is also made to pounding. Drift criteria are set at either
the elastic design force level or the maximum expected displacement.

Throughout the

documents it is assumed that the maximum expected displacements can be obtained as the
elastic displacements computed under design loads times the behavior factor. The ECCS

-69states that the spacing between adjacent structural units shall not be smaller than the sum
of the maximum expected displacements computed for the units.
Drift criteria for damage control differ significantly between ECCS and CEB, but are very
stringent in both cases.

The ECCS states that the interstory deflections

under a

"scnicr:ability earthquake" (not defined) shall not exceed either 0.003h or 0.006h, depending
on the type of nonstructural elements attached to the structure. The CEB limits the elastic
interstory design deflections under specified design loads to (0.01/K)h.

4. 8. 2 (e) Other Relevant Considerations.

Ground motion spectrum is modified for sloping terrain.


Two horizontal and one vertical component of motion have to be considered.
Usually the vertical component
components are combined as

is neglected

A~

and the two horizontal

, with A = 0 for regular structures,

and A = 0.3 otherwise (ECCS).


Static lateral force analysis is permitted only for regular structures with h <
80 m and T < 2 seconds.
Severe penalties are given to irregular structures.
P-o effects have to be considered explicitly whenever PO/(Vh) > O.I, where
0 is the maximum expected story drift, i.e.' 0 = q oe .

4. 8. 3 Evaluation of Ductility Considerations. The global design is similar to, that proposed
in the ATC 3-06 document [ I6]. Elastic ground motion response spectra form the basis for
design. These spectra are modified for design, principally by increasing the spectral values
for T > T0

This is achieved by making the descending branch of the design spectrum

proportional to I I T 213 rather than I I T. No reasons are given for this modification, but
it probably is to account for multimode effects and to provide larger safety for long period
structures (ATC 3-06 arguments). Design force levels (for strength design) are obtained by
dividing this modified spectrum by behavior factors that may be as high as 8 and as low as

2.

-70-

The behavior factors are period independent and based on judgement.

The implicit

assumption (deduced from the criterion that inelastic displacements are equal to elastic
displacement times behavior factor) is that behavior factors and ductility ratios are identical.
In the CEB the behavior factors are independent of the real strength of the structure, thus,
no consideration is given to the increase in structure strength beyond the strength design
level. In the ECCS the behavior factors can be increased by the ratio of structure strength
to first yielding strength.

As a result, steel structures (ECCS) can be designed for a

considerably smaller force level than RC structures (CEB), and significant differences in
ductility demand have to be expected.

In both the ECCS and the CEB the design is based on explicit strength considerations and
only implicit ductility considerations. As a consequence, the expected ductility demands will
be different for the same type of structures, dependent on structural period and many other
It is well established that the ratio of strength reduction factor (behavior

considerations.

factors) over ductility ratio is not a constant. This ratio is period dependent and is greatly
affected by site conditions, particularly for soft sites. For such sites the use of an elastic
response spectrum for design of structures that are expected to respond inelastically in a
severe earthquake is not appropriate unless site specific strength reduction factors are
considered.

At the element

level the ECCS classifies elements according to behavior factors,

distinguishing between "nondissipative" elements and "dissipative" elements. The former are
"brittle" elements which cannot be counted upon to dissipate energy through inelastic
deformations.

The latter are placed into three classes, A, B, and C, corresponding to

behavior factor q < 6, < 4, and < 2, respectively. Specific detailing requirements are given
for the three classes.

In the ECCS great emphasis is placed on "safety checks" whose purpose it is to force
inelastic deformations into dissipative zones (e.g.," ... dissipative zones must be part of the
beams and not of the columns ... ")and prevent overloads in nondissipative zones. The

"

-71-

latter aspect is represented by very stringent requirements on relative strengths of members


and connections.

For instance, for bracing systems it is stated that, ". . . it must be

guaranteed that the axial yielding of the tensile diagonals precedes the ultimate strength of
their connections as well as the collapse of beams and columns." Implementation of this
concept is very rigorous, much more so than in the 1988 UBC, and includes also
consideration of resistance factors (usually 111.2) and material overstrength factors.
The CEB does not contain similar overload safeguards [note the parallel to the 1988 UBC],
but is rather explicit in detailing requirements for individual elements.

4. 8. 4 Evaluation of Drift Considerations.

Drift limitations are more severe than in U.S.

codes, but their intended purpose is equally nontransparent. Specification of drift limitations
at the design force level, which is arrived at by dividing a severe earthquake spectrum by
judgmental behavior factors, cannot provide consistent damage control. It will be difficult
to relate _this drift to that anticipated in a "serviceability" earthquake for which damage
control is intended. The often quoted statement that maximum expected drift in a severe
earthquake is equal to elastic design drift times behavior factor should be looked upon with
great caution.

Much value is placed in this drift, so computed, since it used to specify

building separation and to evaluate

P-~

effects.

4. 8. 5 Summary. The basic concepts of seismic design in both documents are rather similar
to those implemented in modern U.S. codes (ATC 3-06, UBC 1988). Significant differences
to U.S. codes are noted in specified numerical values ofjudgmental design parameters (e.g.,
behavior factors, drift limits) and in design details. Most of the differences point toward
more stringent design requirements for European practice. Both European codes, similarly
to present U.S. codes, are strength based rather than ductility based, and advocate a single
level design without a clear distinction between damage control (serviceability) and collapse
control (safety) requirements.

-73TABLE 4.1 RELATIVE VALUES OF SPECTRUM AMPLIFICATION FACTORS*

Percent of critical
damping

Amplification factor for


Displacement

2.5
2.2

0.5

2.0

1
2
5
7
10
20

Acceleration

Velocity

1.8
i .4
1.2
1.1
1.0

4.0
3.6
3.2
2.8

6.4
5.8
.. 5.2
4.3

1.9

2.6
1.9

1.5
1.3

1.5

1.1

1.2

After Newmark and Han:

Table 4.2 - q FACTORS

Structural
St=~ct......:::.-es

F:=ame

System

\-,
..
mainlv reactincr i:1 .... enclncr:

structures

sa u 1 a...L.

eccentric tr':.lSS bracings

~ 8

braced frame structures


I

Concentric truss bracings


Diagonal bracings

bracings

Cantilever structures
Structures with reinforced concret.e
walls

Mixed structures

a u/ a 1
a u/ a 1

a u/ al

see recommendations
for reinforced
conc:-ete

-75-

f
I

1.2

UBC

1.0

r-' I
I ,

o.s : 1
lt

lt

0.6

IJ

0.4

II

I
l

0.2

0
0

2.0

1.0

3.0

T(sec)

Fig. 4.1. SLEDRS SPECIFIED BY US's UBC AATD JAPAN's BSL

0.4

0.2

0
0

2.0

1.0
T(sec)

Fig. 4.2. JAPAN'S BSL SERVICE LEVEL SLEDRS

3.0

0.6 -

SHEAR WALL
_,JAPAN'S BSL

0.4--- --.........
SMRSF

0.3

SHEAR WALL

..

~/
I

0.18 .,.0.2
__
-0.12~

/'

-SMHSF
--

'

'-.]

'---..

---

--...

--

0\
I

__._us unc

--~

0
0

2.0

1.0

SOIL PROFILE

s.

3.0

(T (sec)

Fig. 4.3- COMPARISON OF JAPAN's BSL AND US's UBC DESIGN EQ's (SIDRS) AT SAFETY LEVEL

-77OAKLAND OUTER HARBOR CIIAN.3

~.4

1.2

II

! It

1.0

I
I

'

I
I

I~

''

II

~=0%
~ =2%
~=5%
~ = 10%
~
20%

~=50%

=60%

0.8

FOR~=

5%

0. 6

0.4
0. 2
0. 0

0.0

0. 5

1.0
1.5
2.0
PERIOD (sec)

2.5

3. 0

Fig. 4.4 - COMPARISON OF DE-AMPLIFICATION OF LERS (CY)


DUE TO THE EFFECTS OF~ AND Jl

J
-78OUCTIUTY=S

OUCTIUTY=2
E:LC:HAOAloiSI-IoiOHRAZ

~1-J,!C:~

0.8

c::

z~

0.6

o.a ................
. . .

!lli2.Qlli :.~~
~F!f_!oi~A_R.!\.-.H.A.t. .L_

.
I t~ ~~~
.
....................
i ;:. ..~;..:;..:.:.~:. -.:....

0.6

u
c::::

:J
c...
::::E

0.4

-~

0.4

<
w

Cl

0.2

0.2

o~---r--~~-~-~.~~~--~--r-~~~~+~~--~~~

0.1

10

0~.--~--~~~~---T~~~~~--~

25

0.1

10

25

FREQUENCY, Hz

FREQUENCY, Hz
Comparison of deamplification fac:ors for 2% damping.

OUCTIUTY=S

OUCTIUTY=2
ELC:HAOAio.ISI-IoiOHRAZ

~1-._!C:~

RIOOELL-N(WI.IARK

c:
0
,.....
u

. .
~

,.,,::

::

.u... ~.:. Jfr,

:
: ; : : ::

- - - - - - - -

:;:

NE:WI.IARK-HALL

. :;;.-:

/,

E:LC:HAOAI.ISI-I.IOHRAZ

0.8

.................

!,!1-..,!C:~

RIOOELL-N(WI.IARK

; ::

:-~.--:

......

: ; :::

-----------:::::

~ lgrrj~_R.!\.-.~.L~--1-T: j :1;

;:

~,

0.6

t?u

c::

:J

c..

::::E

<
w

Cl

0.2

: ~ .
0~.--~--~~~~--~~-T~~r.---~.

10

25

0.1

FREQUENCY, Hz

I=REQUENCY, Hz

Comparison of deamplification factors for 5% damping.

Fig. 4.5 COMPARISON OF DE-AMPLIFICATION FACTORS FOR


DIFFERENT VALUES OF~ AND J.L [41]

10

25

-791.0
0.9

0.8

C"'
I'll

VI

....

Damping

0.7
0.6

Transition Zone
to 10% Damping

5%

I'll
L.
CIJ

o.s

CIJ

u
u

<

0.4

I'll

....
u

Demand

Pt.

L.

0.3

iO%

~Damping

llJ

c.
VI

0.2

Pt. A

Capacity~

0. I

o.s

I
1.0

---I

1.5

-"-1

i.S

2.0

Period, T (sec.)

Fig. 4.6 - CAPACTIY SPECfRUM METHOD [20]

0.8

SPECTRAL
ACCELERATION

Sa(g)

0.6 -:I
/

0.5

SLEDRS

= 5%

~\
IMPULSIVE
EQ

IDRS
Sl

= 1.5, ~ = 6%

r-HARMONIC
EQ

0.4

0.2

-----

0.1 _:'

0.5

1.0

T(secs)

1.5

2.0

Fig. 4.7- RECOMMENDED CAPACI'IT SPECfRUM MEfHOD

2.5

-80-

i vg
.

GROUND MOTION

/""'\

rv u

\.._)

v C\U /'\ . .

1 RESISTANCE

1v RESPONSE OF THE STRUCTIJRE

HYSTERETIC BEHAVIOR
RESISTANCE-DEFORMATION
RELATIONSHIP
(a) HARMONIC 1YPE OF EQ

'

..

vg

GROUND MOTION
RESISTANCE

RESPONSE OF THE STRUCfURE


t

RESISTANCE-DEFORMATION
RELATIONSHIP
(b) IMPULSIVE 1YPE OF EQ

Fig. 4.8 COMPARISON OF STRUCI1JRAL RESPONSE TO DIFFERENT lYPES OF EQ's

-81Rigid and ince:mecii3.ce subsoils


F1exible subsoils

020

I
I
I

I
I

I
I

I
I

Cl-7

Q-1

CoefficientQ-10

-....:.;:"

Zone c.;

I
I

1--

Zon~

'

o-os

I
I

I
I

I
I
I
I

.... I .

i a;s . . a-c
I
...

Zone A

015

I
I

I
I

I
I

02

r.....
---

.,..._

--

1
I
I

I
I
I

I
T
T
0-4

I
I

I
I

~"-

....

......

r-

..,..,

.......

----

"'"'-;

~.J.

.-.:::::

I
I

r......

I """-t-- T

- --oc:s
-075
-065

I
I

-ous
-os

. I

12 ..

t-o

0
in seconds

'!-

I
I

I
!'-..

T
06

Period

I
I

I
I

1'- I
:::-::j,

I4-co

Fig. 4.9- NEW ZEALAND NZ 4203, 1984: BASIC


SEISMIC COEFFICIENT C.

SLEDRS: Ce
0.8

SIDRS: Cu
ZONE A

0.6
FLEXIBLE SUBSOILS
0.4

i
0.2

'-....

--

----"----

0.33}
0.3

..:::_- - - - - - - - - - 0.026 }
0.024

cu

II-

....... ' -

I
0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

(T (sec)

Fig. 4.10 NEW ZEALAND NZS 4203, 1984: SLEDRS (CJ AND SIDRS (Cu)

-82-

Cy

ELASTIC DESIGN SPECTRA (SLEDRS)

1.2 ~--------------------------------------~

usc

l.O

1988

JAPAN-SSL

------------- ..........
...

0. 8

0. 6

NZS-4203
------,
.

...

...

0.4

-...
.. .

... .. ..

. ...
... ...

-------------~-~~~:--------

MEXICO DF-1987
, ~----------- ---

o. 2
'

.........

----------------------------

0. 0

o.s

0. 0

l.O

1.5

2.0

2.5

3.0

PERIOD (sec)
INELASTIC DESIGN SPECTRA (SIDRS)
0 4

0 3

JAPAN-BSL

......................... .

0.2

usc
0. l

.. .

1988

NZS-4203

. ..

... ... ...

....

...... ..... .. .
...

M EX! C 0 D F 198 7 .. - - - .. - - - - - - --~_:-:_:-:_-:-_:-:_-=-=-=-=--=---:-...--c...:....:~

-------------

---------------------------

0 0
0. 0

05

1.0

1.5

2.

2.5

PERIOD (sec)
Fig. 4.11 - COMPARISON OF SLEDRS AND SIDRS

30

-83STIFFNESS AND IDI REQUIREMENTS

0.'

SOIL TYPE 1
SHORT PERIOD
JAPAN-BSL

0. J

...-.::-,;;o~;,;;~ llrnll)

...
......

....
0.2

::

. ......./

o.1

1988 usc

/ ./1984 NZS
~-~t~ r----------' ,,... ..
,~
1987 MEX OF

I
I

...
...
..
.....

', !l,:. . .----------------,.""

:;f./
0. 0

0. 0 0 s

0. 0

0. 0 10

o. a 1s

0. 0 20

0.025

INTERSTORY DRIFT INDEX

Cy

STIFFNESS AND IDI REQUIREMENTS

0.16

. 0. 14

SOIL TYPE 1
LONG PERIOD

0.12

/.:.::(,;c;n;Ci;~um.umt>

JAPANBSL

.:
0. 10

/ .................

0. 0 B

0. 0 6

0. 0 4

./1984 NZS
'---~---~--------:
.
. r...~.-~,9=8~8~U~a=c---

a.o2

------i98"7""M-Ex-i5i="

.'
0. 0

0. 0

s
'INTERSTORY DRIFT INDEX

0. 0 0 s

0. 0 10

0. 0 l

0. 0 2 0

o. a 2s

Fig. 4.12 - CODE EXJ>ECTED BASE SHEAR - IDI RELATIONSHIPS

1
-84a Sa
V
----c-a- g - s-w

0:1 (1985
0.40 1--------,_.;....-..;.
_ _'"'

0.30
01 (1976)

3
2
T (Seconds)

0.0

Fig. 4.13 - COMPARISON OF REDUCED DESIGN SPECfA (DUE TO


DUCTILITY Q) FOR BUILDINGS LOCATED IN GROUP B
LOCATED IN ZONE III OF FEDERAL DISTRICT OF
MEXICO: 1976 AND 1985 (EMERGENCY) CODES

Oakland Harbor
1.2

DEG

- - - J.L
--------- J.L
----- J.L
------ J.L
----- J.L
---- J.L

1.0
0.8

0.6

0.4

I,

',/

'

= 1
= 2

= 3
= 4
= 5

= 6

"
....... ',,.,

-~-,,~~'' -- ,.

0.2

35

,.

....... -

., ..
""--=:: . , ,
~

-,

---'' , ..
':.._..... _.,
'--- .....
""...

'

"

'

~-~~-=t-:..: ::-~ ::::':. . . : :'~----- -------- .. -----~...... ...-::.-..........:;..:.:.=.;,.;:::_ ~!.:----

0.0
0.0

0.5

l.5
2.0
Period (sec)

1.0

2.5

Fig. 4.14 - REDUCTION OF mE IRS DUE TO J.L,

3.0

-85-

A
0

'

~-.:--ID

dOdd 0

Fig. 4.15 - SMOOTHED LINEAR ELASTIC DESIGN RESPONSE FOR A

=5

RCT)!

R <TJ .1.

Ro

2
0-3R

----------

To

1l

A
j

.0

dO

co

Fig. 4.16 - DERIVATION OF THE RECOMMENDED SMOOTHED LINEAR


ELASTIC DESIGN RESPONSE SPECTRA

-875. SUMMARY OF STATE-OF-THE-PRACTICE

ON DUCTILITY

AND DRIFT-BASED EARTIIQUAKE-RESISTANT

DESIGN

5. 1 GENERAL REMARKS
Usually the practice in EQRD is based on official code procedures. Thus, the state-of-thepractice on J.L 6 and IDI-based EQRD has been judged by analyzing and comparing the
present EQRD codes of the United States, Japan, New Zealand, Mexico and Europe (CEB).
In judging the results of the analyses conducted and particularly the observations made from
the comparisons of the codes, it is necessary to keep in mind the following facts:

1.

The studies presented herein analyzed and compared the code regulations regarding
seismic forces and stiffness requirements.

The strength, stiffness and the seismic

behavior of a constructed building in general is not the result solely of specified


seismic forces and minimum stiffness (or maximum acceptable IDI), but it is governed
by the overall design philosophy and the consequence of complex combinations of this
philosophy and the specified forces, and stiffness with: (a) the satisfaction of code
material requirements (such as the minimum requirements for sizing and detailing
structural and nonstructural components); (b) the construction technology [quality
control of materials,

workmanship, and compliance

with design requirements

(drawings), which are related to the degree and quality of inspection]; and (c) the
maintenance

or preservation

of the whole soil-foundation-superstructure-

and-

nonstructural system.

2.

The seismic forces specified in the code reflect the seismicity of the country and this
seismicity varies not only from one country to another, but within countries it also
varies from one region to another.

For example, the severity (damage potential) of

earthquake ground motions in the regions of highest seismic risk in New Zealand is
smaller than the severity of earthquake ground motions in the zones of high seismic
risk in Japan or the United States.

Therefore, the adopted SLEDRS in codes of

different countries cannot be the same. From inspection of the SLEDRS shown in

-88Fig. 5.1, it is clear that the seismic risks in Mexico City and in the worst zone of New
Zealand are significantly lower than those corresponding to the region of highest
seismic risk in Japan and the United States. From the point of view of comparing how
ductility is used to reduce the required linear elastic seismic forces, the ideal would
have been to normalize all the SLEDRS to a spectral acceleration of l.Og.

5. 2 CODE SPECIFIED SLEDRS


5. 2. 1 Sites with Firm Soils (Soil Type 1). From the SLEDRS presented and illustrated
in Fig. 5.1 (a) , it is clear that, except for buildjngs with T < 0.4 sees. and T > 2.1 sees., the
Japanese BSL is the most severe code. For T < 0.4 sees., the UBC is the most stringent,
whereas forT > 2.1 sees., the New Zealand Standards (NZS) code is the most severe. The
smallest elastic forces are required by the Mexican D.F. code, which is understandable
considering the seismicity and geology at and around the Federal District of Mexico.

It can then be concluded that, for tall buildings with T

2.0 sees. and located on firm soil,

the United States and Japan have similar required SLEDRS which are somewhat smaller
(up to 20% for T = 3.0 sees.) than the NZS.

5. 2. 2 Sites with Soft Soils (Soil Type 3). From the analysis of the SLEDRS plotted in Fig.
5.1 (b), it can be seen that, except for buildings with a T < 0.6 sees., the Japanese BSL is
the most severe spectrum. For T < 0.6 sees. the UBC is the most stringent. If the special
soil Profile S4 introduced in the 1988 UBC is also considered, then the UBC becomes as
severe as the Japanese BSL and becomes even more severe forT > 2.0 sees. It should also
be noted that the Mexican D.F. SLEDRS it is as severe as the Japanese BSL forT > 3.2
sees. Then for tall buildings with T > 2.0 sees. and up to T

=4 sees. located on a very soft

soil (soft clay, UBC type S4 or Zone III of Mexico City), the UBC is the most severe
SLEDRS. For these tall buildings the CEB SLEDRS is the least demanding.

-89-

5. 3 USE OF /-, TO REDUCE SLEDRS TO SIDRS


5. 3. 1 Firm Soil (Soil Twe ll. The SIDRS are obtained from the SLEDRS by reducing
(dividing) them by a factor that depends on the 1- 6 The values of 1- 6 vary according to the
type of structure.

For the case of reinforced concrete SMRSF, the resulting SIDRS are

plotted in Fig. 5.1 (c). Comparing these SIDRS with the SLEDRS [Fig. 5.1 (a)] from which
they were obtained, it becomes clear that the largest reductions are those of the UBC,
which is particularly true for buildings with T < 1.5 sees. For T > 1.5 sees., the minimum
specified seismic coefficient specified by UBC

(c/~ ~

0.075) controls the reduction. UBC

uses a constant reduction of 12/1.4 = 8.6. The Japanese BSL uses the smallest reduction:

1.0/0.3 = 3.3. For tall buildings with T > 1.5 sees. and on firm soil the most severe SIDRS
is the Japanese BSL, which, up toT = 3.0 sees., is more than 33% higher than any one of
the other SIDRS.

5. 3. 2 Sites with Soft Soils (Soil Type 3) The SIDRS at the yield level for building on this
type of soil are plotted in Fig. 5.1 (b).

Comparison of the SIDRS with the respective

SLEDRS of Fig. 5.1 (d) allows evaluation of the deamplification that is involved going from
the SLEDRS to the SIDRS.

The smallest deamplification is again that specified by the

Japanese BSL and amounts to a constant reduction using a coefficient 3.33 = 1/0.3. The
largest reduction is that recommended by UBC, which is 12/1.4 = 8.6. From the SIDRS
shown in Fig. 5.1 (d), it is clear that for tall buildings with a T > 1. 7 sees. and up to a T
= 3.0 sees., the yielding strength required by BSL exceeds by more than 30%, 82%, and

121% those specified by the Mexican D.F.,CEB, and NZS codes, respectively. The yielding
strength required by UBC for tall buildings having T > 2.0 sees. is the lowest one of all the
codes considered herein. Even when the UBC Special Soil Profile S4 is considered, still the
corresponding SIDRS require significantly smaller yielding strength than that of the
Japanese BSL, Mexican D.F.,and CEB regulations.

-905. 4 STATE-OF-TilE-PRACTICE

ON THE USE OF IDI LIMITATIONS IN EQRD

Although all of the EQRD codes reviewed herein have regulations limiting the maximum
IDI for certain limit states, none of these codes have recommendations regarding how the
specified limitations should be directly introduced into the preliminary EQRD of a building
structure.

Already in 1977, Ref. 5 suggested a method in which acceptable limits of IDI

were introduced at the preliminary phase of EQRD, however, in present practice (to the
best of the authors' knowledge) the IDI limits specified by codes are checked by analysis of
the already finished preliminary design of the structure. As will be discussed later in more
detail, there is a need to develop practical code procedures that directly include the control
of damage in the initial phase of the preliminary EQRD.

How can this be done? -- By

imposing as one of the main design parameters first a minimum lateral stiffness or a
maximum acceptable IDI at service level and then a maximum acceptable IDI at the safety
(collapse) level. To have an idea of how the different codes attempt to control the damage,
it is convenient to plot the load-deformation diagrams according to the requirements of
these codes. This has been done in Fig. 5.2. To be able to discuss more specifically the
problem of how the maximum acceptable IDI can affect the design of tall buildings, the
code expected load-deformation, or more specifically, the base shear IDI diagrams for the
different codes considered (analyzed) herein have been plotted for two different groups of
T: short T (T < 0.3 sees.); and long T (T > 1.6 sees.).

5. 4. 1 Minimum Lateral Stiffness and Acce.ptable Limits on IDI at Serviceabilityy Level.

5. 4. 1 (a) Short T: From the results plotted in Fig. 5.2 (a), it can be seen that there is a
very large variation in what is considered to be minimum adequate lateral stiffness (or
maximum acceptable IDI at service level). Although, except for the Japanese BSL, none
of the codes considered herein clearly defines what is the strength that is required by service
EQ ground motions, it is clear that the New Zealand Standards is the code requiring the
largest lateral stiffness, and, therefore the one that will result in better damage control
under their service EQs. This is specifically true in the case where

-91nonstructural elements can be damaged (because of the way that they are connected to the
structure). In this case the NZS require the IDI be smaller than 0.0006 which is near 1/2,

1/4, and 1/6 of those specified by CEB, BSL, and UBC, respectively. In the case of Mexico
D.F., the code requirements for the minimum lateral stiffness become very severe for the
case of buildings located in Zone III (Soil profile S3 , i.e. soft soil [Fig. 5.2 (c)]). For this
case the lateral stiffness required is practically the same as that required by the NZS. Why
are there such large differences?

The only logical explanation

would be that the

nonstructural elements used in these countries are different: while New Zealand is using
very stiff partitions and claddings, the United States uses very flexible nonstructural
elements. However the authors believe that at present the differences in the stiffnesses 'and
damageability of these nonstructural elements cannot, on their own, justify the difference
in the maximum acceptable IDI specified by the codes.

These differences should be

investigated further.

5. 4. 1 (b) Long T For building with T > 1.6 sees. and on finn soils (Soil Profile S), the
code expected strength-IDI diagrams are plotted in Fig. 5.2 (b). From this figure it can be
seen that the results regarding the minimum required lateral stiffness (or maximum
acceptable IDI limits) are similar to those discussed above for the case of Short T. In the
case of tall buildings located on soft soils [Fig. 5.2 (d)], the Mexican code becomes as severe
as the NZS regarding the lateral stiffness requirements.

From the above analyses it appears that while the Japanese BSL is the one requiring the
largest strength for service EQ ground motions, as far as lateral stiffness is concerned, these
laws are the most relaxed of all the codes considered, particularly for the case in which the
nonstructural components cannot be damaged.

From inspection of damages observed in

EQs, it would appear that under service EQs, the IDI should not exceed 0.003 when
nonstructural components can undergo the same amount of deformation; however, this limit
needs further research.

-92-

5. 4. 2 Maximum Acceptable IDI at IDtimate Limit States (Collapse). (Fig. 5.2). While
some of the codes explicitly specify the maximum acceptable lateral displacement (in the
case of Mexico D.F.), there are others that limit the IDI in an implicit way (in the case of
UBC). The Japanese BSL does not specify any limit, however, from discussions with the
Kajima research team it appears that in practice the Japanese designers limit the maximum
acceptable IDI to 0.01. The Mexico D.F. explicitly specifies that the maximum IDI should
not exceed the values of 0.006 and 0.012, depending on whether the nonstructural
components can or cannot be damaged. The UBC implicitly specifies that the IDI shall not
exceed the values of 1.5% in the case of buildings with Short T (less than 65 ft. in height)
and 1.125% in the case of buildings greater in height.

The ATC 3-06 recommended

different values for the maximum acceptable IDI, depending on Seismic Hazard Exposure
Groups (SHEG).

For essential facilities (SHEG III), the IDI limit is 0.01 which is

significantly more stringent than the acceptable limits for the other two SHEG which is
0.015. However, these ATC recommendations clarified that for the case of buildings in the
SHEG I, i.e. structures of ordinary importance, the requirement of 0.015 can be relaxed by
as much as 113, resulting in an IDI of 0.020,provided that their height is less than 3 stories
and that there are no brittle-type finishes.

The authors would like to emphasize that the above limits of 1.0% and 1.5% are a
consensus judgment from experience collected from observations and analyses conducted
on what occurred in previous earthquakes. Compliance with these limits will insure not only
human safety but also damage control. From experience gained in previous earthquakes,
it can be concluded that structures designed and constructed in accordance with present
UBC will in general be capable of undergoing lateral deformations significantly higher than
those corresponding to an IDI limit of 0.015 without collapsing, however the damage will
be very high.

Examples supporting the above conclusions are the measured lateral

deformations in: the main building of the Olive View Hospital during the San Fernando
earthquake

of 1971; the Imperial County Building during the 1979 Imperial Valley

earthquake; and the 21 story building in the Pino Suarez complex during the 1985 Mexico
City earthquake (although one of the 21 story buildings collapsed, the other remained

-93standing with IDI larger than 1.5% ). It should also be noted that while the authors agree
with the UBC specification limiting the IDI to 0.0125, this limit should be connected with
a minimum required yielding strength. Un-fortunately, the minimum UBC required strength
is believed to be too low, thus design of tall buildings that attempt to provide the building
structure with just this minimum strength in case of severe EQ ground motions will undergo
IDI significantly larger than the maximum acceptable by the code.

For sites located in

regions of high seismic risk, it would appear that while the minimum requirements for
yielding strength imposed by the BSL are the more realistic, the initial stiffness seems low,
suggesting the need for these laws to limit the maximum IDI.

-95(a) ELASTIC DESIGN SPECTRA

(b) ELASTIC DESiGN SPECTRA

SOIL TYPE 1

SOIL TYPE 3

l.O

1. 0

1987 CES

a.a

0. I

0.1

p9~~-l'!~~\

.
0.(

....
'':.:_..

..

0.'

-....--------
......

1987 MEX OF

....... _._

.-----'-------

..;"'

-------------

o.o ~------------------------------------~
0. 0

0.5

~-------.\
!
;~.
I

,:,:.....

0.2

1954 NZS \

0. (

--~

1987 CE3
,-----...

\.

1. 0

LO

l.

'\

.--'-~7 MEX D_F_ ___

...~'

.....

0.2

---------

.....

.
0. 0

o.o

05

1.0

PERIOD (sec)

(c)

0. J

(d)

INELASTIC DESIGN SPECTRA

SOIL TYPE 1
JAPANSSL

---.

SOIL TYPE :3
JAPANBSL

0 .l

---------...
..

1967

..

,:::: ~z:v~-------.~.~:--
1987 MEX OF

:::: ~:~:: .

0.2

c:::

~'---., __. .. ...


0. 1

l.O

:. 0

a.'

..
0. 2

1.5

PERIOD (sec)

INELASTIC DESIGN SPECTRA

a.'

....._----- ----------------.

"'.: ... :-..:--

~,.

-.

1967 MEX OF

a.1

---. _____

----~-=-=-~=--~:~==

.-:.-:.-:.~

--------------------------------

a.o ~----------------------------------~
o. a
a.s
1.0
l.S
LO
l.O

a.o
a.o

a.5

l. 0

PERIOD (sec)

Fig. 5.1 - COMPARISON OF SLEDRS AND SIDRS

-1. 5

PERIOD (sec)

2. 0

2.5

l.a

(h) STIHI~ESS AND IDIIlJ:QUIIlEMENJS

(a) STIFFNESS AND IDI REQUIREMENTS


0.1 . - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - -

SOil. TYPE 1
SIIOIH PEiliOD

J/\1' AN-OSl

o. I

.................... :.:::=" ............ j;;.;. ;~~;;;~~... 1\ .~,,

.......

.......

a. u
a.

II

a.

ll

SOil TYPE 1
LONG PEiliOD
JAPANOSI.

/:.;(,;o,;;u~;.;;~.i;"'"'J

...

....
....

a. 1 o

.....
.
....

....
/./

0. 3

........./

1907 CEO

0. 01

/:'

rr~--r---------------

; i ..-/ ......
i i /i . :'

0. a1

.
.
0

... !.99.?..~~~gf....

0. 03

o.a

.o
o.o

a.aos

o .a 10

! i /.' .....

NlS
;'ri ,.,.:'.' /--------.
I

~~L'?,O_I\_tllL

0. I

0.

o IS

0,010

o .ozs

; iJ //./

o. a 10
o. a IS
a.aos
'INTEnSTOI'lY DlliH INDEX

o.o

0,15

1900 UDC

rl.' ~,:

(d)

(c) STIFFNESS AND IDI REQUIIlEMENTS

190~

~.-

:,-,/'~ ..-:: ...........................


.
;,.,, .:
1901 MEX DF

INTHISTORY Olliff INDEX

Cy

........ 1907 C[D

:-rt.~---.,'----------------

a. al

1900 IJOC

0 ,OJO

'.OJ s

STIFFNESS AND IDII'lEQUII'lEMENTS

r------------------------------------,

0 .I

SOIL TYPE J
SIIORT PEI!IOD

a. 1 a

0. 15

o. I o

0. OS

SOIL TYPE J
LONG PERIOD

a. o

o.o

o.a

o.oos

0.010

o.o 15

o.ozs

a.oas

a.ola
a .o IS
a .on
INTI!IlSTOilY DRifT INDEX

INTERSTORY DRifT INDEX

Fig. 5.2- CODE EXPECTED BASE SHEAR- IDI DIAGRAMS

a.aJS

\0
0\
I

-976. RESEARCH, DEVELOPMENT, AND EDUCATIONAL NEEDS

6. 1 GENERAL REMARKS REGARDING TilE NEED FOR IDEAL SOLUTION


From the discussions presented on the state-of-the-art on the use of J. 6 and IDI limits in the
preliminary EQRD it can be concluded that the ideal approach could be formulated as
follows.

GNEN:

1. Function of building.

2. Site of building.

3. General configuration of building, structural layout, and structural system.

REQUIRED:

To attain an efficient (optimum) EQRD of the building.

SOLUTION:

To achieve an efficient final solution, requiring an iterative procedure,


it is necessary to start with an efficient preliminary EQRD. To carry out this
preliminary design, it is necessary to establish reliable design EQs in the form
of smoothed seismic design spectra. To do so, it is necessary to collect and/or
develop the following information (data).

1. Catalog of EQ records at the given site complemented

with records

obtained at sites with similar site topography and local soil (geotechnical)
conditions located in tectonically similar regions and if necessary to generate
them

through

the use of semi-empirical

relationships

and applicable

seismological theory. Deterministic and probabilistic approaches similar to


those used by Rosenblueth and his research associates can be followed [50].
The available data should be classified according to their intensity (damage
potential) and their recurrence period.

-982. Based on information collected above under item 1., to establish the
serviceability design EQ in the form of a SLEDRS

for strength and

displacement (IDI).

3. Based on item 1., to establish the design EQ for the ultimate limit states
considering not only human safety but also control of damage in the forms of
SIDRS for strength Cy (in function of J.L 6) and for maximum acceptable IDI.
It should be noted that the value to be adopted for acceptable J.L 6 and IDI will

depend on the function of the building. For very important essential facilities,
the value of acceptable

J.L 6

might be just one, thus in this case the SIDRS

becomes equal to the SLEDRS corresponding to the most severe EQ ground


motions that can occur at the site.

Once the above information

is collected and/or developed,

the procedure

for the

preliminary EQRD of the structure is similar to that described in Ref. 5. Thus, what is
needed is to develop reliable information described above under items 1, 2 and 3. The
following is a brief discussion of the research and development needs to improve EQRD of
buildings.

6.

RESEARCH

AND

DEVELOPMENT

NEEDS

TO

IMPROVE

THE

ESTABLISHMENT OF SLEDRS

6. 2. 1 SLEDRS for Service Level EQs. As has been previously mentioned, except for the
Japanese BSL, none of the codes reviewed herein specify or recommend the use of design
EQ at service level, i.e. their design procedures are based on the use of only one design EQ
which is established at ultimate limit state. There is an urgent need to develop a reliable
preliminary EQRD procedure based on two-level design, where the following two limit
states are considered:

functional continuation (serviceability), i.e. ability to remain elastic

during ground motion that is expected to occur several times during the life of the structure;
and then survivability and control of damage under a rare but possibly severe (extreme) EQ

-99ground motion.

Although these needs were recognised several years ago [4, 5, 10], they

have recently been emphasized in the U.S. [51, 52]. To be able to develop reliable
procedures such as two-level EQRD, it is necessary to conduct statistical and probabilistic
analyses of data available regarding what can be considered service EQ ground motions,
then to develop reliable SLEDRS that consider the LERS of all these ground motions.

Because reliable measured data on EQ ground motions are scarce, design spectra are
currently formulated using insufficient statistical information. Data from records of severe
ground motions of earthquakes that have occurred during the last seventeen years has
altered the previous statistical base so dramatically that drastic changes in the LEDRS and,
therefore, in the code-specified C, have been required. Examples of such ground motions
are: the 1968 Hachinohe earthquake, the 1971 San Fernando earthquake, the 1978 Sendai
earthquake, the 1979 Imperial Valley earthquake, the 1985 Chilean and Mexico City
earthquakes, the latter being perhaps the most dramatic, the 1986 San Salvador earthquake,
and the 1989 Lorna Prieta earthquake. Until 1971, the recorded NS component of the 1940
El Centro earthquake was considered as representative of extreme EQ ground motion. The
records obtained during the 1971 San Fernando Valley and the 1979 Imperial Valley
earthquakes demonstrated, however, that the damage potential of this El Centro component
was very low compared with that of some of the recorded San Fernando and Imperial Valley
motions.

6. 2. 2 SLEDRS for Survivability For this limit state, it is necessary to collect and/or
develop reliable data regarding the most severe EQ ground motions that can occur at the
site of the building during a certain established period. To achieve this it is necessary first
to answer the following question: "What is the right engineering parameter for defining the
severity of any given (recorded) EQ ground motion?" This question had been analyzed in
the studies reported in Refs. 25 and 26, which conclude that the best parameter to measure
the damage potential of an EQ ground motion seems to be its Energy Input,
expected hysteretic energy,

, or the

. These studies, together with those reported in Ref. 36,

demonstrate very clearly that EQs such as the 1940 El Centro (which is usually used as a

-100-

MCEQ to check the safety of designed structures) have a damage potential to structures (as
measured by its

EJ

that is significantly smaller than that of recently recorded motions. This

is illustrated in Fig. 6.1. Furthermore, as illustrated in Figs. 6.2 and 6.3, the SLEDRS
assumed by the 1985 SEAOC (which is the one used in the 1988 UBC) and ATC 3-06 are
significantly smaller than the LERS corresponding to the recorded ground motions during
the 1985 EQs in Mexico and Chile, and during other EQs such as the 1986 San Salvador
and the 1971 San Fernando EQs. Thus, if such ground motions were to occur in the United
States in the future, the values of the SLEDRS adopted by present United States code
requirements will significantly underestimate the response that could occur. It should be
noted that similar conclusions can be drawn if the LERS of these ground motions are
compared with the SLEDRS spectra adopted by the Japanese BSL as well as any of the
other codes considered herein.

6. 3 RESEARCH AND DEVELOPME.Nf NEEDS TO IMPROVE THE


ESTABLISHMENT OF SIDRS

6. 3. 1 SIDRS for Strength. C,. Ideal Solution: For any given site the ideal solution would
be to derive the SIDRS directly from statistical and probabilistic analyses of the IRS
corresponding to all recorded motions at the selected site or at similar sites located in
tectonically similar regions, and even of records derived through the use of theoretical
consideration.

The IRS for the selected ground motions should be computed taking into

account the different (hysteretic) models according to the expected behavior of the type of
structural system and materials to be used. Procedures and techniques for computing these
IRS have been discussed in detail in Refs. 24 and 27. There are a large number of such IRS
presently available; a few of these IRS illustrating the variation of the resistance coefficient
Cy with the ductility ratio

J.1.

are shown in Fig. 6.4. From analysis of the results presented

in this figure, it can be concluded that:

1. The shape of the IRS (i.e. the variation of Cy with T) varies significantly depending on
the predominant frequency (or period

TJ of the recorded

ground motion which in turn

-101-

depends on the site conditions (soil profile and topography) in which the record was
obtained [compare the IRS for the Pacoima (rock) record with the Hollister (firm
alluvium) and that of the Mexico (very soft clay)].

2. There is significant reduction (deamplification) of the LERS (i.e. for J. = 1) produced


by yielding (J. > 1) for structures with a T coinciding with or very close to the
predominant

period

(TJ of the ground motion.

(The explanation

for this

deamplification has been given in Ref. 24). The longer the T, the larger seems to be
the deamplification.
3. The degree of reduction of the LERS (J. = 1) due to J. > 1 decrease for T #= T, and
tends to zero as the T of the structure tends to zero.

From comparison of the IRS presented in Fig. 6.4 with the SIDRS adopted by codes which
are given in Fig. 5.1, it can be observed that:

1.

For sites on firm or medium stiff soils there are already several recorded ground
motions whose IRS exceeds the SIDRS adopted by the codes reviewed herein, even
the one adopted by the Japanese BSL, which is by far the most conservative one.
These are the cases of IRS for the Chile, Pacoima, Derived Pacoima, Corralitos,
Capitola, and Hollister records. Note that this is true even in cases of J. = 6 which
is very difficult to achieve and/or to justify its use.

2.

For soft soil sites (soil profile 53 or 5 4), particularly with soft clays whose depth exceeds
40 feet, from the IRS corresponding to the recorded ground motions in Mexico City,
it appears that the SIDRS corresponding to the Cr adopted by the codes will be
exceeded, except for the case of the Japanese BSL for low and medium rise buildings
(up to 20 stories), even when a ductility J. = 6 could be developed and used. This
conclusion would be particularly true in the case of soft clay which can resist and

-102therefore transfer peak ground acceleration up to 0.30gto the foundation, which seems
to be the value considered by the Japanese BSL for soft soils.

6. 3. 1 (a) Code Procedures to Determine SIDRS for Cy. As has been discussed earlier, the
SIDRS for Cy specified by codes are obtained by deamplifying the LEDRS through the use
of a reduction factor R which depends on !J. 6 Some of the authors have been analyzing the
rationale for the R code values since they were introduced in the codes. As discussed in
more detail in Refs. 8, 9, 21 and 36, it is very difficult to judge the rationale for the values
recommended

for the values of R that ATC 3-06 and the values of R,.. that SEAOC

(adopted by 1988 UBC) have recommended for reducing the SLEDRS to the recommended
SIDRS. These difficulties arise from the lack of discussion or even any indication of how
these values have been derived and what they are meant to physically represent. In Chapter
4 of the A TC 3-06 Commentary, it is stated that R "is an empirical response reduction factor
intended to account for both damping and the ductility inherent in the structural system at
displacements

great enough to surpass initial yield and approach

the ultimate load

displacement of the structural system." In evaluating this statement, it should be noted that
the SLEDRS selected by ATC is already based on 5% damped LEDRS.

Therefore, the

equivalent viscous damping expected in clean structures should not be significantly greater.
If the values of R and R,.. are assumed to depend only on !J. 6 , then the studies reported in
Refs. 8 and 9 clearly demonstrate that for any selected resistance

function~

damping ratio,

and ductility, the reduction factor varies with the period of the structure, decreasing as T
decreases.

This has been clearly confirmed by the results shown in Fig. 6.5 which, for

different recorded ground motions, give the spectra of R .. = LERS/IRS corresponding to


5% damping. Furthermore, these results show that R are not directly proportional to the
!J..

From the above discussion it appears that the recommendation of a constant value for R
(or R...), i.e. that the value be independent ofT for the structure, cannot be justified solely
on the basis of the ductility built up in a structure. The values recommended for R (or R,.)
appear too high, particularly for structures with a T < T, (the smaller is the TIT, ratio, the

-103smaller is the reduction), if the designer attempts to design the structure with just the
strength required by the code, based on the use of the specified or recommended values of
R (orR,.,). Fortunately, as shown in previous publications [8,9,2l],the resulting code design
generally produces a significant overstrength, and the shorter the T of the structure is, the
larger this overstrength seems to be.

A better explanation of R is given in Chapter 3 of the ATC 3-06 Commentary.

"The

response modification factor, R, and ... have been established considering that structures
generally have additional overstrength capacity, above that whereby the design loads cause
significant yield." The authors believe that this overstrength, OVS, together with built-in
toughness is a "blessing" because structures designed according to presently
specified design seismic forces (UBC or recommended ATC values) would be able to safely
withstand MCEQ shaking. The first "significant effective yielding" of properly designed
(sized), detailed, constructed and maintained highly redundant structures with relatively
short Tis not only considerably higher than that on which the code design is based, but such
structures also have a significant overstrength beyond their first effective yielding. The
resulting overstrength usually totals two to three times the minimum code-specified effective
yield strength.

6. 3. 2 Implications of Recent Research Results with Res.pect to Rationale for R Code


Values In Refs. 8 and 9, the authors have analyzed the implications of the results obtained
in shaking-table experiments on a seven-story RC frame-wall test structure, and after
comparing ATC minimum required design strengths, the design strength used, ATC 5%
damped SLEDRS, 5% damped LERS for shaking table motion and measured strengths, the
actual value ofR could not have been larger than 2.7. Therefore, it was also concluded that

"it is very difficult to rationalize (justify) quantitatively the values recommended by ATC for
R." If the value of R alone is used in the design of reinforced concrete frame-wall dual

systems, i.e. without any other requirements, the resulting design will not be reliable. The
use of a specific value for R should be tied to other requirements.

In the present ATC

recommendations, the value of R is tied to stringent requirements for detailing reinforced

-104concrete ductile moment-resisting space frame members and structural walls. The authors
believe that this is not enough, and suggest that the preliminary design using an ATerecommended approach (or that of the 1988 UBC) be subjected to a limit analysis to obtain
an estimate of the actual maximum resistance of the structure as it will be constructed, and
that a value approximately three to five (depending on the structural type and fundamental
period T) times the minimum yielding strength required by ATC be ensured. Furthermore,
the design of walls (sizing and detailing) against shear (as well as against members of ductile
moment-resisting space frames against shear) should be based on their maximum resistance.

Figure 6.6 [25] clearly shows that structures with T


according to ATC 3-06 (J..L 6

= 5.5 and

= 5%)

1.5 sec. which have been designed

will be required to have a yielding strength

(represented by Cy) significantly higher than that required by ATC (represented by C.) or
a strength capacity [defined in the figure as Overstrength Factor (OVF)Rcquircd

= required

Cy I ATC's C.] significantly higher than that required by the ATC 3-06 provisions.
Reference 25 shows that a structure with T

1.0 sec. designed and constructed to just satisfy

the minimum required resistance (C.) by the ATC 3-06 provisions is required to develop
ductility displacement ratios, J.L, well beyond the value usually considered as acceptable (J.L

= 5).
All the above results have been confirmed by the studies that are being conducted on the
recorded ground motions during the 1989 Lorna Prieta EQ [36]. The plot presented in Fig.
6. 7 can also be interpreted as giving an indication of the overstrength needed beyond the
first significant yielding if the design is conducted by using 1988 UBC. For example, if we
assume a RC-SMRSF, the R corresponding to first significant yielding, Ry, rather than the
UBC specified R,., , would be R.)1.4

= 12/1.4 = 8.6.

Thus, drawing the line Ry

= 8.6 on

the plots ofR .. of Fig. 6.5,as it is done in Fig. 6.8,the differences between this line and the

R,. (corresponding to the J.L considered) can be considered as representing the required
overstrength (OVS)Rc<juircd.

This is demonstrated by the graphs of Fig. 6.9. In this figure it

is assumed first that the spectral ordinates are representing the spectral acceleration in
terms of g; thus, they are equivalent to the seismic coefficient C.

Secondly, it is also

-105assumed that the code specified (SLEDRS)cooE is equal to the actual LERS of the critical
EQ ground motion. The reasons why the plots of Fig. 6.8 have just given an approximate
idea and not the accurate representation of the required overstrength follow.

In Fig. 6.9,it is seen that

Rovs

= ~ - R .. )IR .. = (OVS).pc<l,./(SIDRS)cooE

It should be noted

that strictly speaking the (LERS)AcruAL is not the same as the (SLEDR)yeooE . Thus when the
actual LERS of a critical ground motion, or the statistical evaluated (mean or mean plus
one standard deviation) strength demanded for all possible critical ground motions in the
form of a (SLEDRS)AcwAL , differs from the specified code, (SLEDRS)cooE , the use of the
plots of Fig. 6.8 to evaluate

Rovs

= (~. - R.,)/R .. can lead to significant error, as illustrated

in the following examples. First, consider the case of a structure with a T tending to zero
seconds. According to the plots of Fig. 6.8 the

Rovs = (8.6-1)/1 = 7.6. This is not the case,

however, because as T tends to zero the required linear elastic strength, i.e., the
(LERS)AcruAL , tends to the value of the peak ground acceleration as specified by the code
in its (SLEDRS)cooE . In fact, as illustrated in Figs. 6.10 and 6.11, the
= 3.1,and not 8.6. Thus the

Rovs

(~)T-o

1.0/0.32

= (3.1- 1)/1 = 2.1 and not 7.6. Because in general the

code specified (SLEDRS)cooE does not agree with the (LERS)AcruAL or (SLEDRS)AcruAL the
plots of Fig. 6. 8 cannot be used to evaluate required overstrength. Better representation of
required overstrength and of

Rovs are illustrated below for two cases: (a) the case of rock

or firm alluvium sites; and (b) the case of very soft soil sites.

6. 3. 2 (a) The Case of Rock and Firm Soil Sites The expected normalized spectra are
illustrated in Fig. 6.10. From the analysis of the graphs of Fig. 6.10 it appears that, in the
case of structures located on rock or firm alluvium having

J.1.

= 2, the required OVS has the

largest values for structures with T in the range of 0.1 to 0.5 sec. If the structure develops
a

J.1. ~

4, then the largest required OVS seems to occur for aT

=0, 2 sees.

Considering an

effective peak ground acceleration of 0.40 g., i.e., Z = 0.40, the maximum values of OVS
are about

-106-

ovs

~.

= 6, the

OVS

3.6

0.60-0.13=0.47

0.45 -0.13 =0.32

0.4 -0.13=0.27

= 2.5
= 2.1

For T > 0.5 sec. the required OVS decreases.


sees. and p.

Rov

For example, for structures with T = 1.5

= 0.067- 0.042 = 0.025 and

the ROV

= 0.6.

6. 3. 2 (b) Case of Very Soft Soil Sites In the case of deep deposit (larger than 40 feet)
very soft, highly compressible clay, the OVS spectra are quite different from those for rock
and firm alluvium. Fig. 6.11 illustrates the OVS for EQ records on soft soil having a T,
(predominant period of soil) equal to 2 sees. [36].
From Fig. 6.11 it is clear that the normalized specified code (SLEDRS) for values of T
between 0. 8 sees. and 4 sees. underestimates the values obtained from recorded ground
motions. However, as UBC for Zone 4 specifies a Z

= EPA/g = 0.4 no matter what kind

of soil, there appears to be a compensation for the underestimation

in the normalized

ordinates. However, it is clear that for periods around 2 sees. the underestimation is more
than two times. As a result, the EPA is required to be smaller than 0.20 g. As there have
already been. recorded ground motions in such types of soils with EPA significantly higher
than 0.20g. (actually, up to near 0.30g.),it is believed that the code specified spectra when
combined with the Z = 0.4 (which seems to be conservative for soil type S4) is not
conservative for structures with T near the T, of soft soil. From Fig. 6.11 it can be seen that
the required normalized OVS and the

Rovs have the approximate values given in Table 6.1.

Because the code specified (SLEDRS)cooE differs significantly from the (SLEDRS)AcruAL
.. 1.,, ,

the true values

of~

<MEAN

are quite different from the code value of 8.6. Also, as the actual

values of R .. plotted in Figs. 6.11 and 6.8 are obtained dividing the actual values of the
(LERS) by the (IRS) for different values of JJ., it is clear that computation of Rovs from Eq.

I
I

-1076.5 [i.e., Rovs

= (Ry

- R .. )l R .. of Fig. 6.9] using the values of Ry

= 8.6 and

of R .. given by

Fig. 6.8 will result in incorrect values.

There are already many recorded ground motions with (LERS)AcruAL for a certain range of
T which exceeds the (SLEDRS)cooE Examples of these cases are shown in Figs. 6.3,6.7and
6.11. Figures 6. 7 and 6.11 show not only that the spectral ordinates of the code (SLEDRS)
can be significantly exceeded, but that even if a

p. 6 =

6 could be provided and/or used or

accepted (which is doubtful because of the damage that will be involved with such high
displacement ductility ratio) there is a nee.d for significant overstrength beyond the strength
required by Code, particularly for structures with relatively short T, in order to have safe
design. From the results plotted in Fig. 6.5, 6.10, and 6.11, it is clear that the longer the
value of predominant period of the ground motions, T,, the larger the range of the period
of structures for which significant overstrength is required.

From analysis of the records

obtained on soft soils, and particularly those obtained in Mexico City, it would appear that
if very soft clays could resist and therefore induce to the building foundation EQ shaking
with peak acceleration equal to 30% g, the actual IRS will exceed those specified by present
codes, i.e. the (SIDRS)cooE , including the the BSL code for tall buildings.

Lessons learned from analysis of performance of buildings during recent destructive EQs
and results from recent research indiCate that low-rise buildings usually have large
overstrength with respect to that required by the U.S. codes --the taller the building the
smaller the overstrength.

Thus it appears that at least in cities in the United States, the

medium-rise buildings (particularly those located on sites with very soft soils) are the ones
that have to be suspected of becoming a serious threat to life and/or of incurring large
economic loss in the case of a major EQ.

6. 3. 2 (c)

Recommendations

for Improving Code SIDRS~

To improve Code SIDRS

requires improving first the reliability of the specified SLEDRS and then the reliability of
the specified values of R. Therefore, it is recommended:

-1081. To attain more reliable LEDRS

To achieve this it will be necessary to install

appropriate instruments, networks and array to record strong motions in the free field
and at the foundation of structures.

Research should be conducted to improve

processing and probabilistic methods of analyzing strong motion data and quantifying
seismic hazard.
2. To develop more reliable methods for estimating the values of R. This requires a
more precise definition of R. The definition illustrated in Fig. 6.12 is proposed as a
basis for improving the evaluation of R [21, 36].

Once there is sufficient data

regarding possible critical ground motions at any given building site, it will be possible
to obtain the SIDRS directly from the recorded motions, so that there will be no need
to specify the R" . Therefore for the proper use of the definition of Fig. 6.12 in
evaluating reliable values of R, what remains is an urgent need for calibration of the
real strength of structures that have been designed according to present code.

Besides the factors indicated in Fig. 6.12, there is another parameter which can affect R.
One of these extra parameters depends on the soil-structure interaction, which can increase
or decrease the severity of the recorded ground motions at the free field, which are usually
the records considered in the determination of the SLEDRS as well as the SIDRS. Another
parameter is the change in the period of vibration, T, with the degree of damage: as

J.L

increases, the values ofT will increase, and might affect the response.

It has to be noted that all of the above recommendations have been made to improve the

use of J.L 6 in defining reliable SIDRS for preliminary design. This improvement is necessary
but not sufficient.

The damage and failure of a building does not depend only on the

strength given to the structure and the intensity of the ground motions. It also depends on
the number of cycles of inelastic deformations that the structure is required to undergo
during its response, particularly the number of inelastic deformation reversals, which
depends on the duration of strong motion. Thus, it is necessary to define a way to include
the effects of duration of strong motions. These effects will be studied in Task #2 of the
overall research.

Recently conducted studies show that the selected acceptable J.L 6 should

-109depend on the duration of the strong motions. The most promising method for including
the effects of duration of strong motions is through the use of an energy approach.

6. 3. 2 (d) SIDRS for Lateral Displacement and IDI. As has been previously discussed,
practical and simplified methods have been developed to estimate lateral drift [33 - 35].
These methods are based on the assumption that except for buildings with T 1 I T,

< 1, the

nonlinear lateral displacements are similar to the linear lateral displacements. From studies
conducted by the authors [36], it appears that the nonlinear displacements are very sensitive
to the dynamic characteristics of the ground motions and of the structure, and they can be
significantly different than those obtained based on linear behavior as it is illustrated in Figs.
3.5 and 6.13. Note from this last figure that in the case of ground motions with long T,, as
those occurring in very soft soils (the case of Mexico SCT record), the nonlinear
displacement can be significantly smaller (nearly 50% smaller) than the linear displacement
for structures with T 1 close to the value of the T 8 On the other hand, for values T < 2/3T,

== 1.4 sec., the nonlinear displacements are significantly higher (in some cases 3 times
higher). Thus, the authors believe that the ideal solution for finding reliable SIDRS for
lateral displacement is to estimate the IRS for displacement through time history nonlinear
analysis for all the possible types of critical ground motions that can occur at the selected
site and for different SDOF models and then derive a SIDRS through statistical and
probabilistic analyses of these IRS.

Based on the derived IRS, it is possible to estimate lower and upper bound IRS for the
interstory drift for Multi-Degree-of-Freedom Systems (MDOFS). A lower bound is obtained
by assuming constant IDI through the entire height of the structure, and an upper bound can
be estimated by assuming a soft story [25, 36]. Lower bounds of IDis -IRS for different
recorded ground motions and considering

= 5% are given in Fig. 6.14. For any given

building site, once all the possible critical EQ ground motions have been selected and their
IRS for the IDI have been obtained, it is possible through statistical and probabilistic
analyses to derive a lower bound SIDRS for the IDI of MDOFS. This SIDRS can be used
for the preliminary design of a MDOFS.

However, it will be necessary to estimate an

-110amplification factor by which the SIDRS' lower bound should be multiplied to take into
consideration the fact that in the response of MDOFS it is very difficult to achieve a
constant IDI through the whole height of the building. The taller the building (i.e. the
longer the T), the larger the amplification factor should be. Thus there is an urgent need

to conduct integrated analytical and experimental studies to find an empirical expression for

this amplification factor.

TABLE 6.1 VALUES OF NORMALIZED OVS AND OF Rovs COMPUTED FROM FIG. 6. 11

NORMALIZED OVS AND (Rovs) FOR

J.L=2

(sees)

0.9

J.L=4

1.55 - 0.32
1.23
(1.23/0.32 = 3.84)

1.10 - 0.32
(0. 78/0.32

= 0.78
= 2.43)

J.L=6
-

-----

0.80 - 0.32
0.58
(0.58/0.32 = 1.81)

-I

2.0
3.0

1.40 - 0.10
(1.22/1.18

= 1.22

' 0. 75 - 0.18
6. 77) . (0.57/0.18

=
0.80 - 0.14 = 0.66
(0.66/0.14 = 4.71)

= 0.57

= 3.16)
0.50 - 0.14 = 0.36
(0.36/0.14 = 2.57)

= 0.32
= 1.78)
0.45 - 0.14 = 0.31
(0.31/0.14 = 2.21)
0.50 - 0.18
(0.32/0.18

Edm (inch 2 /sec2 )

E1 /m (inch 2 /sec2 )

Chile

20000

Mexico

20000
u2

----

u4

---------

15000

u2

uG

15000

10000

10000

5000

5000

-----

u4

--------

uaG

0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0

0.5

Period (.sec)

1.0

2.0

1.5

2.5

3.0

Period (.sec)
I

El Centro

......
......

San Salvador

VJ
I

20000

20000

15000

15000

uG

10000

10000

5000

5000

....

o.o

0.5

. --

1.0

L5

2.0

2.5

3.0

o.o

0.5

....~

...

1.0

1.5

.......

~.

2.0

Period (sec)

Period (.sec)

Fig 6.1 - INPUT ENERGY SPECTRA

2.5

3.0

-114-

2.4

2.0

LINEAR ELASTIC RESPONSE


SPECTRA FOR ~=5%
& FOR THE 1985 EQs:

-CHILE

--MEXICO

( N 1 OE COMPONENT
AT LLOLLEO}

(EW COMPONENT
AT SCT)

1.6

0.4

0.0

2
T {Seconds)

Fig. 6.2 - COMPARISON OF 5% DAMPED


LERS AND SLEDRS

Spa (g)

3.0

,,,,"
I I

; :__Pacoima
I

2. 5 -

(5% Damping)

,,'

I
I

.,

II

II
II

2.0-

1'

II

II

~~~

'I I
I I I

\'I

I
I I

I
I'

'v''
I,# '
I

San Salvador

1.5
I
,_..

,_..

Ul

1.0

0.8

0.5

0. 0
0.0

0.5

1.0

1.5
2.0
Period (second)

2.5

Fig. 6.3 - COMPARISON OF PSEUDO- ACCELERATION SPECTRA, Spa

3.0

-116-

Deri.ved Paco:i.ma

Mi.yagi.

l.2

L2
ul

u-2

l.O

o.a

I
I

0.4

0.2

u-4
uS

,,'

o.e

u-'

1
1

,' , ...\

0.,

....

,'

'.12

l.O

ul

..;?-~~~~~:::,;.~==

o.o

o.c
0.2

o.o
o.o

0.!

l.O

l.!

PecJ.oc

c,.

2.:1

2.0

3.0

0.0

0.5

Oakland OuLer Harbor Chan. 1

1.2

2.0

1.0
PerJ.oQ

(secl

-~-'

- - - - p-2
- - ~-J

--,.-2
--,.-s
-,.

1.0

- - - p-4

0.1

0.1

--- ,_,

-- p-!t

o.c

3.0

Oakland Outer Harbor Chan. 3

---p-1

1.0

2.5

(secl

o.c

0.4

o.c

0.2

o.:z

o.o

o.a
o.o

0 .5

1.0

1.!5

J.O

2.!5

l.O

a.o

0.!5

Pec.l.od (sec I

Cy

1.0

2 .5

2.0

1.-'

J.O

Period (SeC I

Cy

Oalhnd 7-Story B!dt;. Chan. 7

_,_,

l.:Z

1.0

----- p-2
---p-l

1.0

0.1

___ ,_,

0.1

1.::1

_,_.

Bid~.

__
,_,
----,.-2

Chan. 3

_,_.,_,

--pl

-- p-5

o.c

Oalkand 7-St.ory

--
_,_,

o.c

0.4

o.c

o.:z

O.l

o.o

o.o
o.a

0.!5

1. 0

1.!5

Pec.l.od (SeC I

:z.o

:z.,

J.O

o.o

0.!5

1.0

:.a

1.!5

Pedod <cl

Fig. 6.4- VARIATION OF RESISTANCE COEFF1CIENT WITH 1-'

(~=5%)

2.!5

l.O

-117-

Cy

Ch.ile
1.2

1.2

l.O

1.0

o.a

o.a

0.6

0.6

0.4

0.4

0.2

0.2

0.0

Mex.ico
ul
u2
ul
u4

...,
uS

0.0

o.o

0.5

l.O

1.5

Per.1ocl

Cy

2.0

2.5

3.0

0.0

0.5

(sec)

l.O

2.0

Per.1ocl

Cy

El Centro

1.2

San

2.5

3.0

2.5

3.0

(sec)

Salvador

1.2
ul
u2
uJ
u4
u5

l.O

l.O

... ,

o.a

o.a
0.6
0.4
0.2

0.0

0.0
0.0

0.5

l.O

1.5

Per.1ocl

2.0

2. 5.

3.0

0.0

0.5

(secJ

1.0

1.5

Per.1od

Cy

Paco.ima
1.2

1.2

l.O

1.0

o.a

o.a

0.6

0.,

0.4

0.4

0.2

0.2

2.0

(sec)

Taft
ul
u2
ul
u4
u5

... ,

0.0

0.0
0.0

0.5

l.O

1.5

Per1ocl

2.0
(secJ

2.5

3.0

0.0

0.5

1.0

Fig. 6.4 VARIATION OF RESISTANCE COEFFICIENT WITH

2.0

1.5

Per.1od

(secJ

J.L (~

= 5%)

2.5

3.0

..

,
-118-

Hollister

0 DEG

0.8

---- J.L = 2
---- J.L ... 3
------ J.L = 4

0.6

- - - J.L = 6

0.4

0.(

0.2

O.J

o.o
o.o

o.s

9 0 D E:G

--.u=l

l. 0

---p.-6

0.6

Capitola

l.:l

---p. = 1
-- J.L = 2
- - - J.L .... 3
------ J.L - 4
----p. .... 5

l.O

Cy

---.u = 5

o.o
l.O

l.S

2.0
Period (5ec:)

Corralitos

2.5

0.6

1.5
2.0
Period (sec:)

1. 0

Santa Cru:z;

=a

=
=
....
....
- - - J.L ....

o.a

o.s

0.0

DE:G

- - - p.
- p.
- - - - p.
------ p.
---p.

l.O

3.0

J.S

3. 0

90 DEG

---.u=l

1
2
3
4

=2
=3
= 4.
---J.L = 5

l.O

5
6

o.a

-- J.L
- - - J.L
----- J.L

0.6

- - - J.L

0. 4

0,(

0,2

0.2

0.0

0.0
0,0

o.s

1.0
l.S
2.0
Period (sec)

J.S

o.s

o.o

3.0

Emeryville

l.S

2.0
Period (sec)

l.O

350 DE:G

- - - J.L =
------ J.L =
- - - J.L =
----- J.L-

l.. 0

o.a

1
2
3

4
---- J.L = 5
--J.L ... 6

0.6
0.4

0.2

o.o
o.o

o.s

l.. 0

1.5

::.o

2.5

3.0

Period (sec)

Fig. 6.4- VARIATION OF RESISTANCE COEFFICIENT Wimp. (~=5%)

2.5

3.0

RJL

1940

EL

NS

C~NT!tO

- - - 1-' = 2
------ 1-' = 4
----- 1-' = 6

12
10

,.

.'

I ,, \
,', , ,

...., - .._,....,

-,

,a. ;: I :... ,, ,.
; '-:
.. './

(.

..~

J.1"""2

i i

.'\.-.\

....

90 DE:G

10

"

Capitola

RJ.L

14

.,.

......... - .............. . . .

! : ~\
I

! ... i
! :. ,

... ...'.: ..
....

~ I

...,.

1::::1 . .

---- J.L

1::1

,, '

.. :'i .:,....,,
... .:--~
.....................
.
....

~.

,'

J.L

! ~ ,., _.'\

_,

. \ \ .' . . ,:

, , ,
-"'

'

............

! ,., \

,,
\
! ."

......

0. 5

0.0

1.0

2.0

1.5

0. 0

3.0

2.5

2.0
1.0
1.5
Period (sec)

0.5

Period (:sec)

~.s

3.0

-I

1985 MEXICO

RJ.L

EW

SC~

Santa Cruz

90 DEG

\0
I

14
12

---JL=2
. Jl = 4
----- f.L

10

_,,_

'

-~

'

J.L-2

..

.~

.~

'.

' ...... ',,,

.. .. \
. .. ...
.... ...,
..

.' ,

J.L
----- f.!

'

Cl ..

1::1

,4

I I

,. \/\' ' .,
::-.:;
; i ,,

',I

,
,
,
-. '-._ ......

.:;-._.,..... __..... ,,

...........

2
0

0
00

o.s

1.0

1.5

2.0

2.5

3.0

o.o

0.5

1.0

1.5

2.0

Period (sec)

Period (:sec)

Fig. 6.5- VARIATION OF R" WITH Jl AND T

2.5

3.0

llolllster

RJ.L

DEG

10

Emeryville

,,

, '

,,.,..

'

.. ~'

'

\\

\\

~- ~

:
; .:
,.,.I
.....

. .

.1'!

,..'""'',

! \ ... ,. ,_,
,...
'
;' !
... ..

..... ...,, ... ,41'

....,
..
I

'

......\

- ....,,

....................... .

-- J.l = 4
'\ ----- J.l = 6

~'. \

..
\ '\

--p=2

. J.L=4
,.,., ----- J.l = 6

... ... .'

10

--J.L=2

J50 DEG

.'..,--',,----,

0.0

0. 5

2.0
1.0
1.5
Period (sec)

2.5

0.5

0. 0

J.O

1.0
::z.o
1.5
Period (sec)

2.5

J.O

.....
I

Corralitos

0 DEG

Oakland llarbor

- - - J.L = 2

........ fl = 4
----- J.l = 6

,. ,

J.L=2
J.l = 4 _, .......... .
i i
I
,
,. i ---- J.L = 6 ;
,
b.
_i
\
I
I \ 1 ,,
.. \~.
.j : ,
!~:
I
' -..... ...... ........
;, :
~.
;; ..

'
.
. ! .
.
'... .... :
1'.,-,

. ..
.
. ..... .

0I

JS DEG

. .....

1.,,i ....:

0.0

0. 5

2.0
1.0
1.5
Period (sec)

2. 5

J.O

i':

I.

0. 0

0.

Fig. 6.5- VARIATION OF R" WITH J.l AND T

1.0
1.5
2.0
Period (sec)

2.5

J.O

Resistance Coef.

c;

1. 0 , - - - - - - - - - - - - - - . .
-Chile

0.8

L O r - - - -- - -Pcol
------,

-- t 1 Cent r o
...! ......

Resistance Coef.

Hlco

0.8
-----Tart

0.6
0.6

ATC (R 8, Cd 5. 5)

0.4

- - - - Hlya9l

0.4
0.2
0.2

. - ........

0. 0
0.0

0.5

1.0

1.5

2.0

.. 2.5

3.0

o. o L2:::::~;;;;:;;:~~
0.0

Period (sec)

0.5

1.0

1.5

2.0

2.5

3.0

Period (sec)

....
....N
I

OVF!,(.t)

6.-------------.
-Chile

OVF(IIf'd)

Centro

Pcol"

----Derived Pacol1u
-----Taft

- - - - Hlya91

2
0 i.LL<'.LLJ.:LU:'.L.Lt:.ZLL:LL.t:."LLL.'LLLLL.'LALLLL:LJ

0.0

0.5

1.0

1.5

2.0

Period (sec)

2.5

3.0

o.o

0.5

1.0

1.5

2.0

2.5

Period (sec)

Fig. 6.6 REQUIRED SEISMIC COEFFICIENT (Cy) AND OVERSTRENGTH FACTOR (OVF) FOR SDOFS
DESIGNED IN ACCORDANCE WITH THE ATC FOR S1 AND ASSUMING 1-' = 5.5 AND ~ 5%

3.0

-122CORRALITOS

CllAN.

c=s%
2.0

1988 UBC (SLEDRS)


FOR SO~ TYPE 1

1.5

1.0

IS88 UBC (!DRS)


0.5

--------
o.o
0.0

0.5

1.0
l.S
2.0
PERIOD (sec)
HOLLISTER

CIIAN.

2.5

J.O

c=s%
1.2

1188 UBC (SLEDrtS)


FOR SOIL TYPE 3

l.O
0.8

1988 UBC (IDns)


0.6

---

0.<(

0.2
0.0

o.o

o.s

1.0
1.5
2.0
PERIOD (sec)

2.5

3.0

Fig. 6.7 COMPARISON OF 5% DAMPED IRS OF RECORDS OBTAINED


DURING THE 1989 LOMA PRIETA WITH THE 1988 UBC SLEDRS AND SIDRS

-123-

liS

--.--1.! ==

!2

10

---f.l.-6

Capitcla

90 D&C

ooooooooJ.,; 21

806

=~

----- fJ.

lO

R..,..

--.u""

=~,

806

,,

/ f\ \

---- p. -

/.\\ / ! \'' ',,/\

0'\..,0.~;/-./ '\..ooooo/to<:::~o-.
--<::::::
::z

o.o

1.5
:.a
?ed.cci (sec)

:z.s

l.Q

0.5

3.0

3.a

0.0
Pericci {aec)

R..,.

Sa~t&

'0 D&C

Cru:

117

:.'o,o-._.................. ___ ...,,.

:z
0

o.a

a.5

R"

Uollister

10

0 DEC

806

o.o

3.0

1.5
Pez:icci {sec:)

0.5

l.O
:z.o
Period {sac)

&maryville

R"'

l.S

3.0

lSO DEC

10

--.u- 2

\ \

~---~-------""-- --...... fJ. .. 4

--0--IJ. .. 6

':.

"

\.\
\ ",

.... '.

...-.......---......""'.
---

-~:

OoO

0.5

1.0
loO
Pericd (sac)

C:or:~lito

10

l.5

o.o

J.O

0.5

l.O

1.5

l.O

J. 0

Period (sac)

R..,.

D!:C

10

8.6
--.u- 2
!--==--------......... fl. .. 4

Oakland uarbor

J5

DEC

8.6

----.u-s

0. 0

1.0
lo 0
l.5
Pe:icd (sac:)

J.O

0.0

0 5
0

1.0

1.5

;.o

:.s

J.O

Fig. 6.8 VARIATION OF RJl AND OVERSTRENGm ru(8.6- RJl)Wim T AND J.t

-124PROBLEM:

To Find Relationships Among Different Strength Coefficients Spectra.

GIVEN:

1. A SDOFS with an elastic-perfect plastic mechanical behavior.


2. The Critical Ground Motion, CGM, and its 5% Damped (LERS)cGM as
well as its 5% Damped (IRS)MCGM for a specified JJ..

REQUIRED: Assuming that the SDOFS has been designed on the basis of a
(SIDRS)yeoDE , find relationship between the reduction factor due to
overstrength, Rovs , and the overall design reduction factor, Ry , and the
reduction factor corresponding to the energy dissipated by plastic
deformation, RSL

LERS, IRS, C, (A;ez)

Defining

R~'

and R 5 as follows:

(LERS)ccM=R~'[IRS)ccM=R,.JRs(SlDRS),coDE] ( 6 l)
Then as overstrength spectra,

= (IRS)cGM - (SIDRS)ycODE

(OVS)spectra is (OVS)spectra
and defining

OJ'S

Defining

(OVS)~crra

(SIDRS)

R,

yCODE

- - (6 2)
-Rs 1

(LERS)CGM (6.3)
(SlDRS)yCODE

T (sees)

using Eqs. 6. 1 and 6. 2, Eq. 6. 3 can be rewritten as follows:


R = R"(IRS)cGM =R R =R [R
-1] (6.4)
'Y (SIDRS)
.. s .. OJ'S
yCODE

SLEDRS AND SIDRS


therefore
(SLEDRS)coDE ~

,'

Ry CODE (SIDRS)y CODE

Rovs

Ry-R"

(6. 5)

R"

= R, CODE (SIDRS).., CODE

(SIDRS)y CODE

= (SIDRS).,. CODE

1(S-IDRS~~~D"E- - - - - - __ .

(1.4)

Then

T (sees)

Fig. 6. 9 - DEFINITIONS AND RELATIONSHIPS Al\-10NG DIFFERENT REDUCTION


FACTORS

-125-

NORlvfALIZED C
4

ROCK AND FIRM ALLUVIUM SITES

(SLEDRS)
code
soil rype 54

0.0

0.5

1.0

1.5

2.0

2.5

3.0

T (SECONDS)
Fig. 6.10 - CO:MPARISON OF STRENGTH SPECTRA FOR ROCK AND FIRM
ALLUVIUM SITES

NOR.lv4LIZED C
4

VERY SOFT SOIL SITES

(SLEDRS) code

soil type S4

0.0

1.0

2.0

3.0
T (SECONDS)

4.0

5.0

Fig. 6.11 - COl\lPARISON OF NORl\lALIZED STRENGTH SPECTRA FOR


VERY SOFT SITES

6.0

11
fl
v
(LEOS )t , (IRS )l=( Cy
)t , or C5 = W
~ED
c..r
mOXC..I
2 .S

2.0

1.5

0\
I

1.0

0.5

Cs'
0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

Period (Second)

Fig. 6.12- DEFINITION OF TilE RESPONSE MODIFICATION FACTOR


(R = RJ.I X R( X R 5)

-127DISP.[in)

11
~-5%

-p.=l
-------- J.! = 2
---- J.! = 4
----- p. == 6

12
10

DISP.[inj

NS

C:E:N'l:RO

1940

,,

,_ ......-.--

........

90 DEG

---Jl-1

~-5%

----- Jl - 2
---- ;.t.- 4
----- Jl- 6

12
10

CAPITOLA

'

....

-, ...

'

2
0

1.5

l.O

0.5

0.0

2.0

2.5

. 3.0

2.0
1.0
1.5
PERIOD (sec)

0.5

0.0

Period (sec:)

DIS?.[ir.]
50

DISP.[inj

1985 HE:XI:CO SC':

--p.=l

------ p. = 2

40

---- J.!
----- J.!

30

=
=

--.u=-1

12

--- fJ.- 2
---;.t.- 4

0 DEG
~-5%

-----Jl-6

10

4
6

CAPITOLA

l4

3. 0

2. 5

,,

,'
6

20
10

-s%

2
0

0. 0

l.S

l.O

0.5

!.0

2.5

3.t:

0.5

0. 0

?e:-iod (sec:)

DISP.[i;,.j OAKLAND HARBOR


14

--.u=l

12

-;;.

---- iJ -

12

-----.u= 6

10

DISP.finj OAKLAND HARBOR

35 DEG

10

, 'I

2. 5

1.0
1.5
2.0
PERIOD (sec)

3. 0

JOS.Or::G

--J.'=l
----- iJ. = 2
-- --;: = 4
-----.u=o

,,
"\.~-,

.:...! !

.,.-;.-:-. -..:

0
0. 0

0.5

l.O
1.5
2.0
PERIOD (sec:)

2. 5

3. 0

0. 0

0. 5

1.0
2. 0
1.5
PERIOD (sec)

Fig. 6.13- LATERAL DISPLACEMENT SPECTRA [36]

2. 5

J. 0

DISP .[inJ
14

CORRALITOS

10

DISP.(inJ

DEG
~

--J-L=l
......... I' = 2
---- Jl = 1
----- Jl = 6

12

90

= 5%

HOLLISTER

14

_ _ ,, = l

12

- Jt

90

=2

---- Jl = 1

10

, ..

..,a-----

...... ' . ,/'

----- Jl

=6

~-'

.........

. ,. ' ' ... _, , , - ...


..... ,

.,'"

\.

'~

.....

, ;
,' _,
, ,' ........ .
. .
' .

,'

~I

~'

DEG

...

'
~'

2
0

0.0

CORRALITOS

DISP.(inJ
11

2. 5

2. 0
1.5
1.0
PERIOD (sec)

0.5

=2

12

10

----- Jl = 6

---- Jl = (

=5%

14
12

,
,,

10

2.0
1.0
1.5
PERIOD (sec)

0. 5

DISP .(in J

0 DEG

--p=l

- 1-'

3. 0

0
0.0

HOLLISTER

--p-=-1
- J1
2
---" Jl """ 1
p=6
1::1

2. 5

3. 0

0 DEG
"
,'1

;'\ /.

.....

N
00
I

(=a 5%

,
8

.,,,,..

'!

.~::-

0.0

0.5

2.0
1.5
1.0
PERIOD {sec)

2. 5

J. 0

0
0.0

0.5

2. 0
1.5
1.0
PERIOD (sec)

2. 5

J .0

Fig. 6.13- LATERAL DISPLACEMENT SPECTRA [36]

-129-

IDir...

1940

IDir...

NS

EL C!:N'l'RO

CAPITOLA

0.025

0.025

--.u=l
------ ,t..t = 2
---- ,t..t = 4
----- ,t..t = G

0.020

f-~=-1
--- JJ = 2
---- J.l.
----- J.l.

0.02a

0.015

0.015

0.010

0.010

0. 0 OS

0.005

o.o

90 DEG

c:"

=6

o.a
0.0

l.O

IDIL

1.5

2.0
Period (sec)

0.5

1985

KEXICO

2.5

SC~

0.0

3.0

IDIL

EW

--,t..t=l
. -------- ,t..t = 2
---- ,t..t = 4
----- ,t..t = 5

0.015

1.0
1.5
2.0
PERIOD (sec)
CAP ITO!.A

0.025

0. 025

0.020

0.5

0 020

jJ - - - - - jJ-

3.a

0 DEG

{-5%

--JJ-1
------- JJ - 2
----

2.5

4
6

o. 015
a.01a

0.010
0.005

~
.~

.- .-

a.oos
~-S%

0. 0

0. 0
0. 0

o.s

l.O

1.5

?e:i..cc

IDIT
o.o:;:5

2.0

--- j J -

-1

----- J..l

=a

IDir...

JS .OEG

--J..I=i
~-

0.0

:; 0

1.0
1.5
2.a
PERIOD (sec)

{=S%

OAKLAND HARBOR

0.025
0.020
0. 015

a.010

0.010

0.005

305

--JJ=l
- f.!
2
---- ~ = 4
----- J.l. = 6

0.015

0. 0
0. 0

a.5

(sec:)

OAKLAND HA?..:SO?..
---

0.020

1.5

2.5

J.O

DEG

{-s%

0.005
----~

a.o
0.5

1.0
1.5
2.0
PERIOD (sec)

2.5

3. 0

a. a

a. 5

2.0
1.0
1.5
PERIOD (sec)

Fig. 6.14 - IDI SPECTR.\ (5% DAlv!PING) [36]

2.5

J. 0

IDIL

CORRALITOS

0.025

90

..

--p=l
J.1. = 2
--- J.1. - 4
-----p=6

0.020

IDIL

DEG
(=5%

HOLLISTER

0.025

90

DEG

=5%

-p=l
JL = 2
---- JL = 4
----- JL = 6

0.020
0.015

0.015
0.010

0.010

,,

, ' ,~:\:......
I,.'" ....
...,._,
0.005

,;, ...
1

,.~

/-\
-"-

..

, \"\._

,.--------

~....

--------'":..~-:--:.::-

0.005

---------!

0. 0
0. 0

0.0
0. 5

1.0
1.5
2.0
PERIOD (sec)

2.5

3.0

o.o

0.5

1.0
1.5
2.0
PERIOD (sec)

2.5

3.0

....
I

IDIL

CORRALITOS

0.025

0. 020

IDIL

0 DEG

( =5%

--J.l=-1
J.1. = 2

0.025
0.020

----- J.1. = 6

0.015

0.015

0.010

0.010

0.005

0.005

0.0

0.0
0. 5

0 DEG

-s%

-p=l
J.1. = 2

--- Jt = 4

0. 0

HOLLISTER

1.0

1.5

2.0

2.5

3.0

--- JL = 4

-----I'= 6

0.0

0.5

PERIOD (sec)
Fig. 6.14. IDI' SPECfRA (5% DAMPING) [36]

1.0

1.5.

2.0

PERIOD (sec)

2.5

3.0

-131-

7. SUMMARY, CONCLUSIONS AND RECOMMENDATIONS

7.1 SUMMARY

The state-of-the-practice and of-the-art in the use of the concepts of deformation, ductility,
ductility ratio, drift, and interstory drift indices for attaining efficient Earthquake Resistant
Design (EQRD) of buildings structures have been reviewed.

After discussing the advantages of using an energy approach for the EQRD of structures,
and pointing out the differences between deformation, ductility and ductility ratio, the needs
for providing structures with the largest ductility economically feasible and for controlling
the interstory drift index were discussed in detail. The need for establishing more reliable
design criteria for EQRD of structures was also discussed.

The state-of-the-practice and of-the-art ofEQRD of buildings were reviewed, beginning with
a review of the problems in design and construction of EQ-resistant structures, followed by
a review of present Building Seismic Codes, with emphasis on how the concepts of Ductility
Displacement Ratio, p,, and Interstory Drift Index, IDI, are used, and how they could be
used, to improve the state-of-the-practice according to present knowledge.

The review

covers the building seismic codes of the U.S.,Japan, New Zealand, and Europe (ECCS and
CEB).

Based on a review of the problems encountered in the design and construction of EQresistant buildings, research, development and educational

needs to improve present

knowledge and particularly state-of-the-practice were formulated.

7. 2 CONCLUSIONS

From the studies conducted and the results presented in this report, the following main
observations can be made regarding the use of ductility and drift limits in EQRD:

-132-

Although the advantages of using plastic deformations of the structural material to


dissipate part of the seismic Energy Input

(EJ

to the structure and the need for

limiting the lateral interstory drift have been recognized in the literature, their
implementation, particularly their reliable quantification, has not been accomplished
fully in present seismic design codes;

While it is possible to use the concept of ductility in a vague manner in discussing the
philosophy of ductility-based design, when such philosophy has to be applied in the
EQRD of structures the philosophy has to be quantified, and it is therefore necessary
to use unambiguous parameters;
Although displacement ductility factors, J1,, provide good indications of structural
damage, they usually do not adequately

reflect the damage to non-structural

components. To produce safe and economical structures, seismic design methods must
incorporate drift (damage) control, in addition to lateral displacement ductility, as a
design constraint;

Conventionally computed story drifts may not adequately

reflect the potential

structural and non-structural damage to multistory buildings. A better index is the


tangential story drift index, RT ;

Although the general philosophy of EQRD is well established and is in complete


concordance

with the concept of comprehensive

design, current code design

methodologies fall short of realizing the objectives of the general philosophy. While
the statement of the general philosophy points out the need to consider three different
limit states (criteria for levels of earthquake, i.e.:

service; damage control or

functional; and safety, or survival), in practice, design is typically only carried out for
one criterion (usually safety), on the assumption that the other two would be satisfied
automatically;

I:

-133The growing concern over the costs of earthquake damages (direct, functional, and
indirect) points out the need that more attention be given to control of serviceability
and functionality, i.e., control of damage;

Achievement of reliable and efficient EQRD requires satisfaction not only of the
criterion for strength and toughness, but also the criteria for deformation

and

repairability.

are

Strength,

toughness,

deformation

control

and

repairability

interrelated and hard to define;

The following three main problematical areas have been identified in the earthquakeresistant design of structures: (1) Establishment of reliable critical earthquake input
(design earthquakes); (2) determination of the demands on the entire soil-foundationsuperstructure and non-structural components system; and (3) prediction of the real
supplies to the building at the- moment that an earthquake strikes;

- While a sound preliminary design and reliable analysis of this design are necessary,
they do not ensure an efficient earthquake-resistant

structure. The seismic response

of a structure depends not only on how it has been designed, but also on how it has
been constructed and maintained (monitored and preserved) up to the moment that
the earthquake occurs. There is a need to improve the construction and maintenance
practices of structures;

There are several sources of uncertainty in code-specified procedures

for the

estimation of demands, which can be grouped into two categories: (1) specified seismic
forces; and (2) methods used to estimate response to these seismic forces;

Strength Demands. For regular buildings up to a certain height (240 ft. in the U.S.),
most of the codes in the world recommend the use of equivalent (static), lateral
seismic forces, which are expressed as a base shear V =(C.,IR)W where:
seismic coefficient equivalent to a SLEDRS

(Smoothed

c. is the

Linear Elastic Design

-134Response Spectra) for acceleration, S/g, and R is the reduction factor. Although in
most codes the value of R is given without any explicit reference

to global

displacement ductility ratio, !J. 6 , these values depend implicitly on !J. 6 ;

Structural response is usually estimated using linear elastic analyses of the effects
induced by the equivalent static forces or by these forces multiplied by load factors,
depending on whether the design will be performed using allowable (service or
working) stress, or the strength (load and resistance factor) design method;

There are few countries in which codes recommend the use of limit analysis and limit
design methods (plastic design methods);

Stiffness and Drift Demands. Most seismic codes address design for lateral stiffness
and for drift at service level. Only a few codes explicitly require that the contributions
of torsion should be considered in estimating the maximum lateral drift, and very few
give any guidelines regarding how to deal with the effect of multicomponents of
seismic excitations.

Few codes give explicit requirements

regarding how to estimate P-a effects.

or recommendations

There is a need for more rational code

procedures for estimating the demands regarding the stability effects at ultimate limit
states;

Strength Supplies. Most of the Reinforced Concrete (RC) EQRD codes require that
the supplied strength be estimated using the strength method, in which the required
strength of critical sections are evaluated as a function of just the minimum specified
strength of the materials, and then reduced by a strength (resistance) factor. There
are a few codes in which the design and detailing of the critical regions of the
structure are based on the probable supplied strength capacity to the members and to
their connections and, therefore, to the entire structure. The state-of-the-practice as
reflected by most present EQRD codes for RC buildings does not appear to include

-135the use of the concept of energy dissipation capacity in a rational and reliable way
through the use of the

~,;

Stiffness, Defonnation and Stability Capacities.

Most of the RC codes give only

empirical expressions to estimate the so-called "effective" or "effective linear elastic


stiffness"; they do not specify how to evaluate the change in,stiffness of the whole soilfoundation-superstructure and non-structural components system induced by increasing
damage. There is a need to develop code procedures that will lead to estimation of
the global defonnation capacity of the structure under not only monotonically
increasing defonnation, but also under generalized (repeated reversal) defonnation.
This should be done based on the supplied local energy dissipation capacity of the
structural members (rotational ductility ratio and degradation with repeated cycles, Le.,
local hysteretic behavior);

Present practice emphasizes the use of strength as the primary criterion for
preliminary EQRD.

While preliminary design based on shear strength could be

justified where serviceability controls, it cannot be accepted in cases where the design
is contrplled by the ul.timate (safety) limit state where plastic deformation is accepted.
At safety limit state (mechanism formation and mechanism movement), base shear is
insensitive to variation of deformation and, therefore, to damage. Although there have
been some proposals to base preliminary design on only lateral damage, i.e.,on only
controlling the interstory drift, a practical method of this type of design has yet to be
developed. A more rational approach is one which not only recognizes the importance
of strength and stiffness (control of deformation), but also recognizes that while these
two factors are strongly interrelated in the case of elastic response, they are less
strongly interrelated

in the case of inelastic response.

To control inelastic

deformation, however, it is necessary to provide the structure with a minimum yielding


strength. Therefore, to achieve an efficient preliminary EQRD there is a need to
consider two requirements simultaneously: the strength, based on the rational use of
~'

(hysteretic energy); and the deformation, based on the limitation of IDI;

}
-136The future of EQRD is an energy approach in which the concept of ~' is used in the
derivation

of IDRS through statistical and probabilistic analyses of the IRS

corresponding to all available recorded or expected critical ground motions at the


building site, and design is conducted using limit design methodology with proper
consideration of the possibility of shakedown phenomena;

For the immediate


recommended:

or very near future the following compromise solution is

Use design forces obtained from SLEDRS reduced by reliable

reduction factor R. The values of R must take into account the reductions due to:
hysteretic behavior (J..L 6 ); changes in damping and in the fundamental
vibration of the whole building system; and the real overstrength.

period of

The R should be

period and site condition dependent;

Ideally, the use of either of the above methods should be complemented with time
history nonlinear dynamic analyses of the response of the preliminarily designed
building system to the predicted Maximum Credible Earthquake ( MCEQ), ground
motions that can occur at the site. If this is not possible, the least that should be
conducted is a static nonlinear analysis of the building under monotonically increasing
lateral loads;

To control damage, it is necessary to control deformations. Control oflnterstory Drift


Index, IDI, at Serviceability Level: Present seismic codes specify acceptable limits of
IDI that vary from 0.0006to 0.006. Although the estimation of IDI at the service level
is usually based on linear elastic analyses, there are many uncertainties regarding the
effective stiffness of the structural members, the deformation of the foundation, and
the contribution of the non-structural components.

Analysis of the deformations

should be based on a realistic 3-D model which considers properly the effect of torsion
under multicomponents of ground motions;

-137Control of IDI at the Safety Limit State.

According to present seismic codes, the

acceptable maximum IDI to control damage varies with the type of structure and its
function, usually varying from 0.01 to 0.03. The IDI spectra demands can be estimated
based on the IDRS for strength for the adopted p. 6

The problem in using these IDI

spectra is in making a reliable estimate of the effective period, T. This is so because


of the difficulties in estimating the effective lateral stiffness. The seismic design codes
are not specific about how to estimate the stiffness of members.
structures, this is a difficult task.

In the case of RC

Although some rules have been formulated for

estimating the lateral stiffness of buildings, the real lateral stiffness varies with the
level of deformation;

Most of the practical methods that have been recommended for designing considering
IDI have been based on the assumption that the nonlinear displacement response is
equal to the linear response spectral values provided that the system has certain
minimum

yielding strength.

Recent

studies have shown that the nonlinear

displacements are very sensitive to the dynamic characteristics of the ground motions,
and in some cases the displacement can be significantly higher than those computed
from a linear elastic response.

Empirical formula have been suggested to estimate

deflection amplification factor Cd defined as the ratio of absolute maximum interstory


displacement to the corresponding value from a linear time history analysis;

Seismic components and their input direction can significantly affect the response of
a torsionally sensitive structural system. Ground components applied at the structural
reference axes may remarkably underestimate the response because the structural
maximum response is dependent on the seismic input direction and its magnitude;

Code Comparison. In judging the results obtained from the comparison of different
codes, it is necessary to keep in mind that it is not enough , just to analyze the code
requirements of the seismic forces and minimum stiffness or maximum acceptable IDI
to be used in the design. The designed structure and the seismic behavior of the

-138-

actual structure are not solely the result of specified seismic forces and IDI, but are
governed by the overall design philosophy and the complex combination of the forces
and IDI with many other factors such as:
requirements;

The satisfaction of code material

the construction technology; and the maintenance or preservation of

the entire soil-foundation-superstructure

and nonstructural

components

system.

Furthermore, the seismic forces in the code of one country reflect the seismicity as
well as the seismic risk of that country, and these factors vary considerably not only
from one country to another, but even from one region to another within a country;

Except for UBC, all the codes reviewed herein consider that portions of the live loads
are seismically reactive and are included in the computation of the seismic forces;
For strength (ultimate or capacity) design there are significant differences in the values
specified by the different codes for the load factors as well as the ways that the loads
are combined;

The codes reviewed herein are strength-based rather than ductility and damage
control-based, and with the exception of the Japanese BSL, advocate a single level
design;

Although the UBC and New Zealand, NZS, code recognize in their material
specification the possibility for overstrength, the only code that explicitly recognizes
and accounts for "maall structural overstrength" due to inelastic redistribution of
forces is the ECCS;

Although most of the seismic codes that have been reviewed permit damage that will
not jeopardize human life, none explicitly defines what constitutes acceptable damage.
Most of the codes recognize that the level of acceptable damage has to be different
for different types of facilities depending

on its occupancy type or function.

Quantitatively, this is accomplished by increasing the seismic forces through an

-139importance or risk-to-life factor. However, the values adopted for this factor seem to
be very low, and it appears to be incompatible with the fact that essential facilities and
those housing very hazardous materials should remain practically elastic. The values
for the occupancy factor, specified by the different codes reviewed herein, varied from
1 to 2;

Code Specified SLEDRS. For buildings with a fundamental period ofT ;;:: 2 sees. and
located on firm soil, the U.S. and Japan have similar required SLEDRS which are
somewhat smaller (up to 20% forT = 3.0 sees.) than the

>

.c.~zs.

For buildings with T

2.0 sees. and up to T=4 sees. located on very soft soil (soft clay, UBC type S4 or

Zone III of Mexico City), the UBC specifies the most severe SLEDRS, and the CEB
has the least demanding SLEDRS;

Use of

~.

to Reduce SLEDRS to SIDRS. All codes except the Mexico Code use a

constant reduction factor, i.e., independent of the T of the structure.


Site with Firm Soil: The largest reductions are those in the UBC. The Japanese BSL
uses the smallest reduction (3.3). The BSL reduction is based on the energy dissipated
only by cracking and local yielding since it does not allow the yielding of the structure
as a whole system (mechanism movement). For tall buildings with T

> 1.5 sees. and

up to T = 3.0 sees., the SIDRS specified by the Japanese BSL is more than 33%
higher than any one of the other SIDRS.
Site with Soft Soils (Type S3

):

The largest reduction is that recommended by UBC

which is 8.6, and the smallest is that specified by the Japanese BSL (3.3). For tall
buildings with a T > 1. 7 sees. and up to T = 3.0 sees., the yielding strength required
by BSL exceeds by more than 30%, 82% and 121% those specified by the Mexican
D.F., CEB and NZS codes respectively. The yielding strength required by UBC for
tall buildings having T > 2.0 sees. is the lowest one of all the codes considered herein;

Use of IDI Limitations in EQRD. Although all of the seismic codes reviewed herein
have regulations limiting the maximum IDI for limit sta.tes, none of these codes have

-140recommendations regarding how the limitations should be directly introduced into the
preliminary EQRD of a building structure.

The IDI limits specified by codes are

checked by analysis of the already finished preliminary design of the structure;

Minimum Lateral Stiffness and Acceptable Limits on IDI at Service2bility Levels.


Short T (T

< 0.3 sees.):

The NZS requires the largest lateral stiffness and therefore,

should result in better damage control under service EQs. This is specifically true in
cases when nonstructural elements can be damaged: IDI :S 0.0006 which is 1/2, 1/4
and 1/6 of those specified by CEB, BSL and UBC respectively.
sees.):

Long T (T

> 1.6

In the case of buildings located on firm soils, the results regarding the

maximum acceptable IDI limits are similar to those for short T. For buildings located
on soft soil, the Mexican D.F. code requirements become as severe as the NZS;

Maximum Acceptable IDI at IDtimate Limit States (Safety).

The Mexico D.F.

explicitly specifies that the maximum IDI should not exceed the values of 0.006 and
0.012 depending upon whether the nonstructural components can or cannot be
damaged.

The UBC implicitly specifies that the IDI shall not exceed the values of

1.5% in the case of buildings less than 65 feet in height and 1.125% for buildings
greater in height. Although the Japanese BSL does not specify any limit for the IDI
at the Safety Level, in practice the Japanese designers limit the IDI to 0.01. These
limits are a consensus judgment from experience based on observations and analyses
conducted during previous Eqs. Compliance with these limits will ensure not only
human safety, but also damage control, provided that these limits are connected with
a minimum required yielding strength. The minimum UBC required yielding strength
seems to be too low. Thus, the design of tall buildings that attempts to provide only
this minimum strength will undergo significantly larger IDI than the maximum
acceptable by the code in case of severe EQ ground motions;

-141Efficient EQRD.

Achieving an efficient EQRD requires an iterative process. It is

necessary to start with an efficient preliminary EQRD. To carry out this preliminary
design, it is necessary first to develop (establish) reliable design Eqs;

There is an urgent need to develop a reliable preliminary EQRD procedure based on


two-level design Eqs, in which the following two limit states are considered:
Functional continuation (serviceability) under frequent ground motions;

and then

survivability and control of damage under a rare but possible severe (extreme) EQ
ground motion;

To enable development of reliable procedures for establishing a two-level EQRD, it


is necessary to conduct statistical and probabilistic analyses of available data regarding
what can be considered service and safety EQ ground motions, and then to develop
reliable SLEDRS and SIDRS that consider the LERS and IDRS, respectively, of all
available recorded or predicted motions in these two levels of EQ ground shakings;

Because reliable measured data on EQ ground motions at different sites (soil profile
and topography) was scarce until 1987, design spectra are currently formulated using
inadequate statistical information.

SIDRS for Strength, C1 For any given site, the ideal solution is to derive the SIDRS
directly from statistical and probabilistic analyses of the IRS corresponding to all
recorded motions at the selected site or at similar sites located in tectonically similar
regions and even of records derived through the use of theoretical considerations;

The shape of the IRS (i.e., the variation of Cr with T) varies significantly depending
on the predominant frequency (or period T,) of the recorded ground motion which
in turn depends on the site conditions (soil profile and topography) from which the
record was obtained;

-142-

There is significant reduction (deamplification) of the LERS (i.e. ,for JJ. = 1) produced
by yielding (JJ. > 1) for structures with a T coinciding with or very close to the
predominant period (TJ of the ground motion. The longer the T,, the larger seems
to be the deamplification;
The degree of reduction of the LERS due to JJ. > 1 decreases as T deviates from T,
and tends to zero as T tends to zero;

Because of the uncertainties in estimating the values ofT, and T, caution should be
taken in applying in practice the observed reduction of the LERS due to
JJ.

> 1 when TIT,

= 1;

For sites on firm or medium stiff soils (types S1 and SJ, there are already several
recorded ground motions whose IRS exceeds the SIDRS adopted by the codes
reviewed herein. This is true even in cases of JJ.

= 6 which

is not only very difficult

to achieve (supply), but also very difficult to justify its possible use because of the
damage that will be involved;

For soft soil sites (soil profile S3 or S4), particularly with soft clays whose depth exceeds
40 ft., from the IRS corresponding to recorded ground motions which can resist and
transfer ground acceleration of 0.30 g to the structure foundation, it appears that the
SIDRS corresponding to the Cy adopted by all codes will be exceeded even when a JJ.

= 6 could

be supplied and used. The only exception is the SIDRS specified by the

Japanese BSL for low and medium-rise buildings of perhaps up to 20 stories;

Code Procedures to Determine SIDRS for C,.. The SIDRS for Cy specified by codes
are obtained by deamplifying (reducing the LERDS) through the use of a reduction
or behavior factor.

Although this factor depends on JJ., it is difficult to judge the

rationale for the values recommended in the codes;

-143The values recommended by the UBC (i.e., R.,) appear too high, particularly for
structures with a T < T, if the designer attempts to design the structure with the
strength required by the code: The value for the reduction factor should be tied to
other requirements besides the value of J.L. The values of the reduction factor should
be affected by the real strength capacity, i.e., the overstrength above the yielding
strength specified by the code;

For structures designed according to UBC, the required overstrength depends on the
J.L,

T, soil conditions and design methodology;

In the case of structures located on rock or firm alluvium, the required normalized
overstrength has the largest values for T in the range of 0.1 to 0.5 sees. and varies
from 0.47 for J.L

= 2 to

0.27 for J.L

= 6.

The corresponding required Reduction for

Overstrength, R...,, varies from 3.6 to 2.1;

In the case of very soft soils, the longer the value of the predominant period of the
ground motions, T,, the larger is the range of the period of the structures, T, for which
significant overstrength is required. The normalized overstrength for a T of 0.9 sees.
can vary from 1.23 for J.L

= 2 to 0.58 for J.L = 6 and

the corresponding Rov. varying from

3.84 to 1.81. The Rov for aT of 2.0 sees. can vary from 6.77 for

J.L

= 2 to

1.78 for

J.L

= 6;

U.S. low-rise buildings usually have large lateral seismic overstrength with respect to
that required by U.S. codes. The taller the building, the smaller the overstrength is.
Thus, it appears that the medium-rise buildings (particularly those located on sites with
very soft soils) are the ones that have to be suspected of becoming a serious threat to
life and/or incurring large economic loss in case of a major EQ.

-1447. 3 RECOMMENDATIONS

7. 3. 1 ReCommendations for Improving Code SIDRS for Strength. C,.

Attain a more reliable LEDRS;

Develop more reliable methods for estimating the values of the reduction factor; This
requires more precise definition of this factor. Although the values of the reduction
factor are affected by several parameters, the main two are the energy dissipated
through hysteretic behavior (damping ratio

and particularly IJ.) and the real

overstrength;

The ideal solution is to attain reliable SIDRS directly from the recorded and/or
analytically derived ground motions. This will eliminate the need for specifying R ...
Therefore, for the proper use of these SIDRS, what remains is to calibrate the real
strength (overstrength) of structures that are designed according to present code;

There is a need to consider in the inelastic design of structures the effects of the
duration of strong motions which include the accumulative ductility and number of
yielding reversals. This can be accomplished through the use of an energy approach
estimating the critical required Hysteretic Energy,

Ea;

There is a need to find reliable factors that will permit the use of the computed
SIDRS for SDOF systems to design MDOF systems;

As it is very difficult to design MDOF structures that will develop uniform story 1J. 6
throughout its height, there is a need to investigate a possible concentration

of

required 1J. 6 at one or more stories and to establish the yielding overstrength required
to limit the maximum
system.

J..1. 6

to the target ductility used in the design based on SDOF

-1457. 3. 2

Recommendations

for Improving SDIRS for Lateral Displacement and IDI.

Nonlinear displacements are very sensitive to the dynamic characteristics of the ground
motions and of the structure, and they can be significantly different from those obtained
based on nonlinear behavior.

For ground motions with long T,, the nonlinear displacement can be significantly
smaller (nearly 50% smaller) than the linear displacement for structures with T
On the other hand, for values T

=T,.

< 2/3T,, the nonlinear displacements are significantly

higher. The smaller the TIT, ratio, the larger the difference is, and it tends to be
proportional to the value of J.L.

Based on derived SIDRS for strength of SDOF systems, formulate SIDRS for
displacement of SDOF systems for different

and J.L.

Based on the derived SIDRS for the displacement of SDOF systems, obtain lower and
upper bounds for the IDI of MDOF systems.

As it is difficult to achieve a constant IDI throughout the entire height of a MDOF


structure, there is an urgent need to investigate (analytically and experimentally)
values of an amplification factor for which the SIDRS' lower bound of SDOFS systems
should be multiplied to obtain a reliable SIDRS for MDOF systems.

-147-

REFERENCES

1.

Housner, G. W., "Limit Design of Structures to Resist Earthquakes," Proceedings,


World Conference on Earthquake Engineering, Berkeley, California, June 1956, pp.
501-511.

2.

Blume, J. A., Newmark, N. M. and Corning, L. H., "Design of Multistory Reinforced


Concrete Buildings for Earthquake Motions," Portland Cement Association, Chicago,
Illinois, 1961.

3.

Anderson, J. C. and Bertero, V. V.,"SeismicBehavior ofMultistory Frames Designed


by Different Philosophies," Report No. UCB/EERC-69111, Earthquake Engineering
Research Center, University of California, Berkeley, California, October 1969.

4.

Bertero, V. V. and Kamil, H., "Nonlinear Seismic Design of Multistory Frames,"


Canadian Journal of Civil Engineering, 2, No. 4, 1974, pp. 494-516.

5.

Zagajeski, S.and Bertero, V. V.,"Computer-Aided Optimum Design ofDuctile RIC


Moment-Resistant

Frame," Proceedings,

Workshop

on

Earthquake-Resistant

Reinforced Concrete Building Construction, University of California, Berkeley,


California, July 1977, III, pp. 1140-1174.

6.

Bertero, V. V., et al., "Earthquake Simulator Tests and Associated Experimental,


Analytical and Correlation Studies of One-Fifth Scale Model," SP-84 American
Concrete Institute, Detroit, Michigan, 1985, pp. 375-424.

7.

Bertero, V. V., "State-of-the-Art in the Seismic Resistant Construction of Structures,"


Proceedings,

Third

International

Microzonation

Conference,

University

Washington, Seattle, Washington, June 28-July 1, 1982, II, pp. 767-808.

of

-1488.

Bertero,

V. V., "Lessons Learned

Implications for Earthquake-Resistant


No. UCB/EERC-86/03,

from Recent

Earthquakes

and Research

Design of Building Structures in U.S.,"Report

Earthquake Engineering Research Center, University of

California, Berkeley, California, April 1986.

9.

Bertero, V. V., "Evaluation of Response Reduction Factors Recommended by ATC


and SEAOC," Proceedings, Third U.S. National

Conference

on Earthquake

Engineering, Charleston, South Carolina, 1986, Earthquake Engineering Research


Institute, El Cerrito, California, III, pp. 1663-1673.

10.

Bertero, V. V., "Earthquake Resistant Reinforced Concrete Building Construction,"


Proceedings, Workshop, I, pp. 38-39, University Extension, University of California,
Berkeley, California, July, 1977.

11.

Mahin, S. A. and Bertero, V. V., "Problems in Establishing and Predicting Ductility


in Aseismic Design," Proceedings, International Symposium on Earthquake Structural
Engineering, St. Louis, Missouri, August 1976.

12.

Naeim, F., "The Seismic Design Handbook," Van Nostrand Reindhold, New York,
1989, Chapter 6, pp. 171-209.

13.

Bertero, V. V. and Bresler, B., "Design and Engineering Decisions: Failure Criteria
(Limit State)," Proceedings, Sixth World Conference on Earthquake Engineering,
New Delhi, India, January 1977.

14.

Dowrick, D. J., "Earthquake Resistant Design for Engineers and Architects," John
Wiley & Sons, Ltd., Sussex, England, 1987.

15.

Cheng, F. Y. and Ger, J. F., "The Effect of Multicomponent Seismic Excitation and
Direction on Response Behavior of 3-D Structures," Proceedings, Fourth U.S.

-149National Conference on Earthquake Engineering, Palm Springs, California, May 2024, 1990, Vol. 2, pp. 3-14.

16.

Applied Technology Council (ATC), "Tentative Provisions for the Development of


Seismic Regulations for Buildings," ATC 3-06, Palo Alto, California, 1978

17.

NEHRP, "Recommended Provisions for the Development of Seismic Regulations for


New Buildings," Federal Emergency Management Agency (FEMA), Building Seismic
Safety Council, Washington D.C., 1988.

18.

Uniform Building Code, International Conference of Building Officials, Whittier,


California, 1988.

19.

American Concrete Institute, Detroit, Michigan, "Building Code Requirements for


Reinforced Concrete," ACI 318-83, 1983.

20.

Department of the Army, the Navy and the Air Force, "Seismic Design Guidelines
for Essential Buildings," February 1986.

21.

Bertero, V. V., "Ductility-Based Structural Design," Proceedings, Ninth World


Conference on Earthquake Engineering, Tokyo-Kyoto, Japan, August 1988, Vol. VIII,
pp. 673-685.

22.

Biggs, J .M., "Introduction to Structural Dynamics," McGraw Hill, Inc., New York,
1964.

23.

Newmark, N. N. and Hall, W. J., "Procedures and Criteria for Earthquake Resistant
Design," Building Practices for Disaster Mitigation, Building Science Series 46, U.S.
Department of Commerce, National Bureau of Standards, Washington, DC, February

-1501973, and "Earthquake Spectra and Design," Earthquake Engineering Research


Institute, El Cerrito, California, 1982.
24.

Bertero, V. V. et al., "Establishment of Design Earthquakes -Evaluation of Present


Methods," Proceedings, International Symposium on Earthquake Structural Methods,
St. Louis, Missouri, August 1976, pp. 551- 580.

25.

Uang, C.-M. and Bertero, V. V., "Implications of Recorded Earthquake Ground


Motions on Seismic Design of Building Structures," Report No. UCB/EERC-88113,
Earthquake

Engineering Research

Center, University of California, Berkeley,

California, November 1988.

26.

Uang, C.-M. and Bertero, V. V., "Use of Energy as a Design Criterion in


Earthquake-Resistant
Engineering

Research

Design," Report

No.

UCB/EERC-88118,

Earthquake

Center, University of California, Berkeley, California,

November 1988.

27.

Mahin, S. A. and Bertero, V. V., "An Evaluation of Inelastic Seismic Design Spectra,"
Journal, Structural Division of the American Society of Civil Engineers, 107,No. ST9,
September 1981, pp. 1177-1195.

28.

Fenwick, R. C. and Davidson, B. J., "Moment Redistribution in Seismic Resistant


Concrete Frames," Proceedings, Pacific Conference on Earthquake Engineering, New
Zealand, August 1987, pp. 95-106.

29.

Cili, F., "Appropriate Yield Strength Distribution for

Low- Rise Reinforced

Concrete

Structure," Proceedings, Eighth European Conference on Earthquake Engineering,


Lisbon, Portugal, September 7-12, 1986, Vol. 5, pp. 8.1/95-108.

-15130.

Otani, S. et al., "Moment Redistribution

in Earthquake

Resistant

Design of

Reinforced Concrete Frames," Proceedings, Ninth World Conference on Earthquake


Engineering, Tokyo-Kyoto, Japan, August 2-9, 1988.
31.

Freeman,

S.A., Czarnki, R.M., and Honda,

K.K, "Significance of Stiffness

Assumptions on Lateral Force Criteria," Reinforced Concrete Structure Subjected to


Wind and Earthquake Forces, Publication SP-63, American Concrete Institute,
Detroit, Michigan, 1980.

32.

MacGregor, J. G. and Hage, S. E., "Stability Analysis and Design of Concrete


Frames," Journal, Structural Division of the American Society of Civil Engineers, No.
STlO, October 1977, pp. 1953-1970.

33.

Sozen, M. A., "Lateral Drift of Reinforced Concrete Structures Subjected to Strong


Ground Motions," University of Illinois, Urbana, Illinois, various pagings.

34.

Shirnazaki, K., "Strong Ground Motion Drift and Base Shear Coefficient for RIC
Structures," Proceedings, Ninth World Conference on Earthquake

Engineering,

Tokyo-Kyoto, Japan, August 2-9, 1988, Vol. V, pp. 165-170.

35.

Shirnazaki, K. and Sozen, M. A., "Seismic Drift of Reinforced Concrete Structures,"


Technical Research Report, Hazarna-Gurni Ltd., 1984, pp. 145-166.

36.

Bertero, V. V. and Miranda, E., "Analysis of the Implications of the Ground Motions
Recorded

during the Lorna Prieta Earthquake," In-House Report, Earthquake

Engineering Research Center, University of California, Berkeley, California, June


1990.

-15237.

Kappos, A. J., "Sensitivity of Calculated Inelastic

S~ismic

Response to Input Motion

Characteristics," Proceedings, Fourth U.S. National Conference on Earthquake


Engineering, May 20-24, 1990, Palm Springs, California, Vol. 2, pp. 25-34.

38.

Hwang, H. and Jaw, J. W., "Statistical Evaluation of Reflection Amplification Factors


for Reinforced Concrete Structures," Proceedings, Fourth U.S. National Conference
on Earthquake Engineering, May 20-24, 1990, Palm Springs, California, Vol. 2, pp.
937-944.

39.

Hatamoto, H. et al., "Seismic Capacity Enhancement of RIC Frames by Means of


Damage

Control Design," Proceedings, Fourth U.S. National

Conference on

Earthquake Engineering, May 20-24, 1990, Palm Springs, California, Vol. 2, pp. 279288.

40.

"Recommended Lateral Force Requirements and Tentative Commentary," Seismology


Committee, Structural Engineers Association of California, San Francisco, California,
1988.

41.

Elghadanisi, F .E., and Mohraz, B., "Inelastic Earthquake

Spectra," Journal of

Earthguake Engineering and Structural Dynamics, 15, pp. 91-104, 1987.

42.

U.S. Army, "Seismic Design Guidelines for Upgrading Existing Buildings," Technical
Manual:

43.

Army TMS-809-10-2, 1986.

Deirlein, G. C. and Hsieh, S.-H., "Seismic Response of Steel Frames with Semirigid
Connections

using the Capacity Spectrum Method," Proceedings, Fourth U.S.

National Conference on Earthquake Engineering, May 20-24, 1990, Palm Springs,


California, Vol. 2, pp. 863-872.

-15344.

Applied Technology

~Council,

"Comparison of Building Seismic Design Practices m

the United States and Japan," Report ATC-15, Palo Alto, California, 1984.

45.

Nakajawa, K. and Johnston, R. C., "Comparative Design of 9-Story Reinforced


Concrete Building using ATC 3-06, UBC 1982 and Current Japanese

Code,"

Proceedings, Ninth World Conference on Earthquake Engineering, August 2-9, 1988,


Tokyo-Kyoto, Japan, Vol. V, pp. 1095-1100.

46.

Okoshi, T. and Merovidi, A. T., "Comparative Design of 10-Story Reinforced


Concrete Building using UBC 1982, ATC 3-06 and Japanese Building Code,"
Proceedings, Ninth World Conference on Earthquake Engineering, August 2-9, 1988,
Tokyo-Kyoto, Japan, Vol. V, pp. 1101-1106.

47.

Teramoto, T. et al., "Comparison of 20-Story Reinforced Concrete Building Design


using ATC 3-06, UBC 1982 and Current Japanese Code," Proceedings, Ninth World
Conference on Earthquake Engineering, August 2-9, 1988, Tokyo-Kyoto, Japan, Vol.

v, pp.
48.

1107-1112.

Tahara, T. and Kamei, T. T., "Comparative Design of 10-Story Steel Building using
ATC 3-06, Los Angeles City Code and Current Japanese Code," Proceedings, Ninth
World Conference on Earthquake Engineering, August 2-9, 1988, Tokyo-Kyoto,
Japan, Vol. V, pp. 1113-1118.

49.

Hisatoku, T. et al., "Comparative Design of 19-Story Steel Building using ATC 3-06,
UBC 1982 and Current Japanese Code," Proceedings, Ninth World Conference on
Earthquake Engineering, August 2-9., 1988, Tokyo-Kyoto, Japan, Vol. V, pp. 11191124.

-15450.

Rosenblueth, E. et al., "The Mexico Earthquake of September 19, 1985 - Design


Spectra for Mexico's Federal District," Earthquake Spectra, Earthquake Engineering
Research Institute, Vol. 5, No. I, February 1989, pp. 273-292.

51.

Whitman, R. V., "Workshop on Ground Motion Parameters for Seismic Hazard


Mapping," July 17 & 18, 1989, Technical Report NCEER 89-0038, National Center
for Earthquake Engineering Research, State University of New York, Buffalo, New
York, December 1, 1989.

52.

Dang, C.-M.and Bertero, V. V., "UBC Seismic Serviceability Regulation: A Critical


Review," under review for publication in the Journal of Structural Engineering,
ASCE.

EVALUATION OF DAMAGE POTENTIAL OF


RECORDED GROUND MOTIONS

Report on Task 2 of the CUREe/Kajima Project on

DESIGN GUIDELINES FOR DUCTILITY AND DRIFT LIMITS

by
Helmut Krawinkler
Aladdin Nassar
Mohsen Rahnama
and
U.S. and Kajima Project Research Team

A CUREe-Kajima Research Report

June 1991

TABLE OF CONTENTS

1. INTRODUCTION

2. DESIGN :METHODOLOGY AND SEISMIC DEMAND PARAME1ERS


2.1 Introduction
2.2 Ductility and Drift Based Design (Dual Level Design)
2. 3 Seismic Demand Parameters

3
3
4

3. SEISMIC DEMANDS FOR SINGLE DEGREE OF FREEDOM SYSTEMS


3.1 Introduction
3.2 Hysteresis Models
3.3 Statistical Study with Firm Soil Records
3.3.1 Ground Motion Records Used in this Study
3. 3. 2 Seismic Demands for Bilinear Systems
3.3.3 Seismic Demands for Stiffness Degrading Systems
3. 4 Damage Potential of Lorna Prieta Ground Motions
3.4.1 Lorna Prieta Ground Motions
3.4.2 Attenuation Characteristics for Rock and Alluvium Sites
3.4.3 Strength Demands for Ground Motions on Soft Soils
3.4.4 Damage Potential of Ground Motions
3.5 Seismic Demands for Selected Records from Japan
3.6 A Pilot Study on Cumulative Damage

31
33

4. SEISMIC DEMANDS FOR MULTI DEGREE OF FREEDOM SYSTEMS


4.1 MDOF Structures Used in this Study

35
35

4.2 Ductility and Base Shear Demands


4.2.1 Story Ductility Demands
4.2.2 Base Shear Demands
4.3 Shear Force and Overturning Moment Demands
4.3.1 Story Shear Demands
4.3.2 Story Overturning Moment Demands

10
16
16
17
20
21
21
23
26

28

39
39
42
47
47
49

5. SUMMARY AND CONCLUSIONS

52

6. ACKNOWLEDGEMENTS

53

REFERENCES

54

TABLES

56

FIGURES

62

EVALUATION OF DAMAGE POTENTIAL OF


RECORDED GROUND MOTIONS
Report on Task 2 of a Project on

DESIGN GUIDELINES FOR DUCTILITY AND DRIFT LIMITS

1.

INTRODUCTION

Seismic design is simply an attempt to assure that strength and deformation capacities
of structures exceed the demands imposed by severe earthquakes with an adequate margin
of safety. This simple statement is difficult to implement because both demands and
capacities are inherently uncertain and dependent on a great number of variables. A
desirable long-range objective of research in earthquake engineering is to provide the basic
knowledge needed to permit an explicit yet simple incorporation of relevant demand and
capacity parameters in the design process. To this end, much work needs to be done.
Identification and evaluation of relevant ground motion and seismic demand parameters is
one aspect of this work. A demand parameter is defined here as a quantity that relates
seismic input (ground motion) to structural response. Thus, it is a response quantity,
obtained by filtering the ground motion through a linear or nonlinear structural filter. A
simple example of a demand parameter is the acceleration response spectrum, which
identifies the strength demand for elastic single degree of freedom (SDO F) systems.
Considering that most structures behave inelastically in a major earthquake, it is evident that
this parameter alone is insufficient to describe seismic demands. Relevant demand
parameters include, but are not limited to, ductility demand, inelastic strength demand, and
cumulative damage parameters such as energy demands.
The term damage potential is used here to denote the potential of ground motions to
inflict damage to manmade structures. This potential depends on the "severity" of ground
shaking as well as the ability of the structure to resist this shaking. Thus, both demands
and capacities need to be considered in assessing the damage potential. In this report the
emphasis is on demand evaluation, but capacity issues are considered in several instances,
particularly in the evaluation of the damage potential of ground motions recorded during the
Lorna Prieta earthquake.

Three sets of ground motions are utilized in this study; a set of 15 ground motions
representative of near-source rock and firm soil motions recorded in Western U.S.
earthquakes, a set of 51 motions recorded during the October 17, 1989 Lorna Prieta
earthquake, and a set of 8 ground motions from Japanese earthquakes provided by Kajima
Corporation. Several seismic demand parameters are evaluated, using various types of
SDOF and MDOF (multi-degree of freedom) models.
The short-range objectives of the work summarized here are to illustrate the feasibility
of assessing seismic demand parameters and damage potential with simplified analytical
models and to evaluate the sensitivity of seismic response to ground motion and structure
characteristics. The long-range objective is to demonstrate that simple yet rational
demand/capacity models can be used to replace the present empirical code design approach
with a more transparent approach based on fundamental principles.

2. DESIGN METHODOLOGY AND SEISMIC DEMAND PARAMETERS

2.1

Introduction
Present code design is based on elastic "strength demand spectra" (e.g., the product

ZCW in the U.S. 1988 UBC), which are scaled to design base shear spectra by means of

system dependent but usually period independent reduction factors (Rw in the 1988 UBC
orR in the ATC-3 approach). The elastic strength demand spectra are modified versions of
smoothened SDOF response spectra, the primary modification being the raising of long
period values to account for multi-mode effects and provide more safety for multistory
structures. Even though designs based on this approach appear to be satisfactory, there are
two conceptual problems with this approach. Firstly, it is well established that the
relationship between system ductilities and R-factors is nonlinear, particularly for short
period structures (data will be shown later). Secondly, because of variable overstrength
that comes from many sources, present designs will result in lateral strengths that vary
widely, dependent on the type of structural system and the number of stories (natural
period). As a consequence, the ductility demands that code designed structures may
experience in a severe earthquake have little relation to the relative R-factors and to the
design base shear, and the design process cannot provide consistent protection for different
structural systems ..
Examples of predicted local (member) ductility demands for three families of code
designed (1988 UBC or SEAOC) steel structures are shown in Fig. 2.1 (Osteraas and
Krawinkler, 1990). All structures have the same plan and story height (72'x120'x12') and
24'x24' bays, and the number of stories is increased till the code period reaches 4 seconds
(except for braced frames which are permitted only to a height of 160', i.e., 13 stories).
All structures are designed by code, but many sources of overstrength are included in the
estimate of structure strength. The local ductilities shown in the figure are estimated from
the strength of the structures, statistical R-Jl-T relationships discussed later, and
relationships between structure and member strengths. Although the true ductilities may
differ from the shown values, two patterns are clearly established. Firstly, the ductility
demands vary significantly with period (i.e., with the number of stories), and secondly, the
braced frames have about the same local ductility demand as the full moment frames (every
frame moment resisting) and perimeter frames, even though they are designed for 50%
more base shear (Rw = 8 versus Rw = 12). This example is shown to document the
arguments made in the previous paragraph.

For the quoted reasons, and others discussed later, there appears to be a need for a
different and more transparent design approach that permits better tuning of the design to
the ductility capacities of different structural systems and the elements that control seismic
behavior. In addition, the issue of damage control deserves special consideration and
should be separated from that of safety against collapse. The following section summarizes
a proposed design approach that forms the motivation for this study on damage potential
and seismic demands imposed by ground motions.

2.2

Ductility and Drift Based Design (Dual Level Design)


The proposed approach is a dual level design approach in which damage control and

safety against collapse are treated as separate design levels associated with earthquakes of
different probabilities of occurrence (damage and collapse threshold earthquakes, see
Osteraas and Krawinkler, 1990). Although the designs for these two limit states are
separated, both designs follow the same basic concept and a reconciliation of generated
design constraints is performed in the process. A flow chart illustrating the proposed dual
level design process is shown in Fig. 2.2.
The basic concept on which both designs are based is an explicit capacity/demand
concept in which both capacity and demand parameters are considered explicitly and the
objective is to design structures in which the capacities exceed the demands with a
sufficient margin of safety. Thus, both capacity and demand parameters must be well
defined, due consideration must be given to the uncertainties inherent in capacities and
demands, and, as a final result, a transparent reliability based design approach needs to be
developed. There is much research work to be done in order to fulfill this objective, and
the study summarized here is nothing but a small step in providing some of the basic
concepts and data needed for this purpose.
This study is not concerned with the issue of damage control. It focuses only on
design for safety against collapse, and there only on the problem of ductility capacities and
demands and the development of strength criteria that result in acceptable capacity/demand
ratios.
In the design for safety against collapse it is postulated that the ductility capacity of

the important lateral load resisting elements is the basis for seismic design. For "brittle"
elements (e.g., many types of connections, vertical load carrying columns that may buckle,
etc.) the ductility capacity is 1.0, and for "ductile" elements it is a quantity to be determined

from expenmental/analytical studies. For the latter elements the ductility capacity is a
function of the deformation history, i.e., of the number, magnitude and cumulative damage
effect of all inelastic excursions, the duration of strong motion, and other cumulative
parameters (Krawinkler et. al., 1983). It is proposed that in the design procedure the
element ductility capacity be modified (weighted) to account for anticipated cumulative
damage effects. This points out the importance of strong motion duration, frequency
content of ground motions, and period of the structural systems, since they all affect the
number and magnitudes of inelastic excursions, which in tum determine the cumulative
damage experienced by a structural component.
In order to derive strength requirements, the element ductility capacities have to be
transformed into story ductility capacities (often a straight forward geometric
transformation), which are then used to derive "inelastic strength demands" for design
(discussed later). The so derived strength demands denote the required strength of the
structure. Accepting the fact that the design profession prefers to perform elastic rather
than plastic design, the structure strength level has to be transformed to a member strength
level in order to perform conventional elastic strength design. Pilot studies have shown
that this transformation is usually not difficult but may require an iteration (Osteraas and
Krawinkler, 1990). After this preliminary design an important step is design verification
through a static incremental nonlinear analysis to verify that the required structure strength
is achieved and to assure that "brittle" elements are not overloaded (ductility less than 1.0).
Clearly, there are many issues in this design approach that have not been addressed
and that may complicate the process considerably. But the proposed approach has been
shown to work in simple examples (Osteraas and Krawinkler, 1990), and deserves further
study to explore its potential. In order to implement the approach, much fundamental
information needs to be developed on the following aspects:
1. Ductility capacities of ductile elements (not part of this study).
2. Statistical data on cumulative damage demand parameters needed to modify (weigh)
ductility capacities.
3. Statistical data on inelastic strength demands for prescribed ductility capacities,
using SDOF systems.
4. Statistical data on the effects of higher modes in MDOF systems, needed to modify
the inelastic strength demands derived from SDOF systems.

In this study several sets of ground motion records are utilized to evaluate the
demand parameters listed in 2. to 4. by performing inelastic time history analysis on
various SDOF and MDOF models. The study is rather comprehensive for near-source firm
soil and rock motions, which are reasonably uniform in their characteristics and permit a
statistical evaluation. For soft soil motions, characteristic patterns are investigated but no
quantitative conclusions can be drawn without a much more comprehensive parameter
study of soil characteristics. The seismic demand parameters used in this study are
summarized in the next section.

2.3

Seismic Demand Parameters

Seismic demands represent the requirements imposed by ground motions on relevant


structural performance parameters. In a local domain this could be the demand on axial
load of a column or the rotation of a plastic hinge in a beam, etc.. Thus, the localized
demands depend on many local and global response characteristics of structures, which
cannot be considered in a study that is concerned with a global evaluation of damage
potential. In this study only SDOF systems and simplified MDOF systems are used as
structural models. Assuming that these models have a reasonably well defined yield
strength, the following basic seismic demand parameters play an important role in
assessing the damage potential of ground motions. Some of the terms used in these
definitions are illustrated in Figs. 2.3 and 2.4

Elastic Strength Demand, Fy,e This parameter defines the yield strength
required of the structural system in order to respond elastically to a ground
motion. For SDOF systems the elastic acceleration response spectra provide the
needed information on this parameter.
Ductility Demand, Jl This parameter is defined as the ratio of maximum
deformation over yield deformation for a system with a yield strength smaller
than the elastic strength demand Fy,e
Inelastic Strength Demand, F y(JJ.). This parameter defines the yield
strength required of an inelastic system in order to limit the ductility demand to a
value of Jl.

Strength Reduction Factor, Ry(J.L).

This parameter defines the reduction

in elastic strength that will result in a ductility demand of J.L. Thus, Ry(J.L) =
Fy.efFy(J.L). This parameter is often denoted as R.
Energy and Cumulative Damage Demands.

Repeated cyclic loading is

known to have a detrimental effect on inelastic response characteristics. There


are many attempts reported in the literature on assessment of cumulative damage
effects through energy terms or specific cumulative damage models. The
energy terms evaluated in this study are
Input Energy, IE:

The energy imparted to the structure by


a ground motion

Damping Energy, DE:

The energy dissipated in the structure


through viscous damping

Hysteretic Energy, HE:

The energy dissipated in the structure


through inelastic deformation

Total Dissipated Energy, TDE:

TDE=DE+HE

Many cumulative damage models have been proposed in the literature, the
simplest one being of the form
N

D = C

(.1Dpi)c

(2.1)

i= 1

where D

cumulative damage

C,c
N

structural performance parameters

number of inelastic excursions experienced in the

.18pi

earthquake
= the plastic deformation range of excursion i (see Fig. 2.4)

For bilinear systems this expression reduces to the total hysteretic energy if the
coefficient Cis taken as the yield strength Fy and the exponent cis taken as 1.0.
For components of steel structures the exponent c was found to be in the range
of 1.5 to 2.0 (Krawinkler, et al., 1983). The coefficient C varies widely and
depends strongly on the performance characteristics of the structure. In the
comparative study performed here, C can be eliminated and the plastic

deformation ranges t18pi can be normalized with respect to the yield deformation

8y. Thus, relative damage can be evaluated from the expression


N

D =

(t18pJ8y)c

(2.2)

i=l

In this study this expression is evaluated using exponents of c = 1.0, 1.5, and

2.0.
The plastic deformation ranges t18pi in Eqs (2.1) and (2.2) are not the ranges as
they appear in the time history response. For the purpose of cumulative damage
evaluation these ranges have to be reordered because small excursions have to
be considered as interruptions of bigger ones. This can be accomplished with
one of several available cycle counting methods. For low-cycle fatigue damage
the rain-flow cycle counting method was found to be the best suited one
(Krawinkler, ei al., 1983). The plastic deformation ranges so identified,
together with the number of inelastic excursions, N, provide basic information
needed for cumulative damage modeling.
There are many other cumulative damage models that have been proposed in the
literature specifically for reinforced concrete structures and elements. Chung et
al., 1987, provides a comprehensive summary of these models. One of these
models is utilized in a pilot study discussed in Section 3.6.
The list of seismic demand parameters enumerated here is by no means complete.
But for conceptual studies much can be learned by using these parameters to assess the
damage potential of ground motions. In the following chapter these parameters are
evaluated for various SDOF systems for closely spaced periods in order to permit a
representation in terms of spectra, using a period range from 0. f sec. to 4.0 sec. In
Chapter 4 the strength and ductility demands are evaluated for three types of MDOF
systems, using six discrete periods covering a range from 0.22 to 2.05 seconds.

3. SEISMIC DEMANDS FOR SINGLE DEGREE OF FREEDOM SYSTEMS

3.1

Introduction
This chapter is concerned with an evaluation of seismic demands imposed by ground

motions on SDOF systems. The use of these simplified models for seismic demand
assessment for real structures is limited, as many important effects present in complex multi
degree of freedom structures cannot be represented adequately. Nevertheless, these simple
models are a very useful tool for conceptual studies, as they provide basic information,
show trends and patterns, and permit comprehensive and statistically acceptable parameter
studies. It is clear' that much of the derived information has to be modified in order to be of
relevance to MDOF structures. Some of these modifications are discussed in Chapter 4.
Iri the context of design, the following questions are addressed in this study.
How can important structural characteristics, including deterioration in
strength and stiffness, be represented in SDOF hysteresis models?
What are representative elastic and inelastic strength demand spectra for firm
soil and rock sites?
What is the relationship between elastic and inelastic strength demand
spectra, which is represented by the strength reduction factor Ry(p).
What consistent patterns exist in parameters that are useful for cumulative
damage evaluation, such as the hysteretic energy dissipation?
How sensitive are the seismic demand parameters to characteristics of the
hysteresis models used to represent structures?
How do seismic demand parameters attenuate with distance from the fault
rupture?
To what extent are the seismic demand parameters affected by site soil
conditions?
How can the damage potential of ground motions be assessed from
simplified SDOF models?

3.2

Hysteresis Models

The SDOF models used in this study are intended to represent global response
characteristics of structures. As such, they should be capable of representing all important
structural characteristics that may significantly affect the response to ground motions. In
addition to basic elastic stiffness and yield strength, the following characteristics deserve
consideration:

Multilinear load - deformation response, representing either the propagation of


cracking or the redistribution of internal forces after yielding of the weakest
element. In the simplest case, such a model should be trilinear.

Strain hardening after formation of a mechanism.

Degradation in stiffness.

Deterioration in strength.

Pinching of hysteresis loops caused by crack opening/closure in reinforced


concrete structures and connection slip or post-buckling behavior in steel
structures.
In order to incorporate these phenomena and study their effect on seismic demand

parameters, a general hysteresis model was developed. The characteristics of this general
model are shown in Fig. 3.1, and the types of specific hysteresis models that can be
generated through rule modifications are summarized in Table 3.1. The specific hysteresis
models that can be represented by the general model are of the following three types.

1.

Nondeteriorating models:
These models are the basic bilinear and trilinear models, which are well established

and need no further elaboration.

2.

Stiffness degradation models with limited memory:


These models exhibit stiffness degradation which is based on memorizing only the

largest positive and negative excursions of past loading cycles. The following two well
known models are included in this category:

10

Peak oriented model, or modified Clough model. In this well known


model the reloading stiffness degrades to the extent that reloading is directed towards the
previous maximum peak point in the direction of loading. This model was proposed by
Clough and was an early attempt to account for pinching of hysteresis loops observed in
experimental studies. From the many modifications proposed for this model in the
literature, the only one incorporated here is that proposed by Mahin and Bertero, 1975,
since it improves the modeling of reloading following small excursions.
Pinching model. A more complex model for the representation of pinching is that
partially illustrated in Fig. 3.1. It is similar to the peak oriented model, with the major
difference being that reloading consists of two steps. Initially, reloading is directed
towards a point defined by the maximum peak displacement and a reduced target resistance
F p The so defined pinching stiffness defines the reloading path until the displacement
attains a value equal to the permanent displacement of the largest previous excursion.
Thereafter, the reloading path is directed towards the previous maximum peak point.
Compared to the peak oriented model this model results in larger pinching, with the amount
of additional pinching determined by the difference between the pinching target resistance
and the peak resistance of all previous excursions.
These two models have two important characteristics in common. Firstly, the
hysteresis loops are identical if several cycles of the same displacement amplitude are
executed. Thus, no further stiffness degradation occurs in these additional cycles.
Secondly, if loading continues beyond the previous peak points, the loading path follows
the originally defined skeleton curve.

Thus, no strength deterioration occurs.

Experimental evidence makes it necessary to refine these models further since history
dependent stiffness degradation and strength deterioration may occur in elements and
complete structures.

3. History dependent strength deterioration and stiffness degradation


models:
It is well established from experimental work and analytical studies that strength and
stiffness properties of elements and structures deteriorate with time. Time implies here the
complete load-displacement history experienced by the system. Materials, and therefore
elements and structures, have a memory and the present state depends on the cumulative
damage effect of all past states. In concept every excursion causes damage, and damage
accumulates as the number of excursions increases. The damage caused by small (elastic)

11

excursions is usually small and probably negligible in the context of seismic behavior.
Thus, only inelastic excursions need to be considered, and from those the large ones cause
significantly more damage than the smaller ones. Also, the cumulative damage may not
lead to a noticeable deterioration until several cycles have been executed This is illustrated
in Fig. 3.2(a), which shows the load-displacement response of a steel beam that failed by
crack propagation aild fracture at a beam-to-column flange weld. Such elements have a
large deterioration threshold, but deteriorate rapidly once deterioration is noticeable, as is
illustrated in Fig. 3.2(b). For other elements deterioration may be evident early in the
loading history (small deterioration threshold) and may occur at an almost constant rate.
The response of such an element is shown in Fig. 3.3(a), and the mode of deterioration is
illustrated in Fig. 3.3(b ).
These examples show that it is necessary to model history dependent deterioration in
a general hysteresis model that is supposed to capture all important phenomena that affect
seismic response.

Deterioration depends on a great many parameters, the simplest ones

being the plastic deformation range (see Fig. 2.4) and the hysteretic energy dissipation in
an excursion. The latter parameter is used here because hysteretic energy is a well defined
quantity in all cases, whereas the plastic deformation range is an ambiguous quantity for
stiffness degrading systems. It is assumed that the hysteretic energy dissipation capacity is
a known quantity and that it is independent of the loading history. The latter assumption is
hard to justify, but it has to be made in order to make a parameter study manageable. The
deterioration per excursion is assumed to be defined by a deterioration parameter f3, given
by the following expression:

(3.1)

where

f3
Ei
Er
LE-}
c

= parameter defining deterioration in each excursion


= hysteretic energy dissipated in excursion i
= hysteretic energy dissipation capacity = aFy8y
= hysteretic energy dissipated in all previous excursions
= exponent defining the rate of deterioration

12

This deterioration parameter is small at the beginning of loading, increases as more


energy is dissipated through inelastic cycles, and approaches infinite as IEj approaches 1
The latter state is associated with complete deterioration defined by zero strength or
stiffness. The rate of deterioration can be controlled through the exponent c. A reasonable
range for this exponent is between 1.0 and 2.0, as suggested for damage modeling of
components of steel structures (Krawinkler et al., 1983). A value of 1.0 will accelerate
early deterioration and slow down deterioration in later cycles, whereas a value of 2.0
achieves the opposite. This is illustrated in the examples given in Figs. 3.4 and 3.5, which
will be discussed later. The energy dissipation capacity 1 depends on the strength and
ductility capacity of the system and can be expressed as the strain energy Fy8y multiplied by
a system dependent parameter a.
The parameter {3 can be utilized to incorporate various types of deterioration
phenomena into the general hysteresis model. At this time, the following three phenomena
are incorporated:

Strength Deterioration:
Deterioration in strength is modeled through a modification of the yield strength F y
and a corresponding translation of the strain hardening stiffness Kh. The modification of
the yield strength is given by
(3.2)

where

= new deteriorated yield strength


Fdo
,
= old deteriorated yield strength (previous value of yield strength)
f3s = (1 - {3) > 0, with the value of f3 given by Eq. (3.1), using the
Fd,n

appropriate parameters to model strength deterioration.


The initial value for F d,o is F Y and the modification is applied after each inelastic
excursion. The parameter f3 determined from Eq. (3.1) cannot exceed 1.0 in this case, or f3s
would become negative. An inspection ofEq. (3.1) shows that a value of {3 > 1.0 will be
obtained only when the system strength approaches zero.
Two examples of the effect of applying this deterioration model to a bilinear
hysteresis model are shown in Figs. 3.4 and 3.5. In both cases the loading history is the
same, consisting of constant amplitude cycles with an amplitude of 48y. The value of Et is

13

assumed to be 100Fy8y (i.e., a= 100), and the exponent cis taken as 1.0 in Fig. 3.4 and
2.0 in Fig. 3.5, respectively. As can be seen, the rate of deterioration is very different in
the two cases because of this difference in c.

Degradation of Unloading Stiffness:


Degradation in the unloading stiffness Ku is modeled in the same manner as strength
deterioration. The modification of the unloading stiffness is given by
Ku,n

where

= (1 - /3) Ku,o = f3u Ku,o

K u,n

= new degraded unloading stiffness

K u,o

= old degraded unloading stiffness (previous value)

f3u

(3.3)

= (1 - /3) > 0, with the value of f3 given by Eq. (3.1), using the

appropriate parameters to model stiffness degradation.


All comments made for strength deterioration apply here as well. The two corresponding
examples illustrating degradation of unloading stiffness are shown in Fig. 3.6.

Accelerated Degradation of Loading Stiffness:


Basic degradation of the loading stiffness is modeled by the previously described
peak oriented or Clough model. This model does not account for stiffness degradation
caused by cumulative loading effects. This additional effect can be superimposed on the
peak oriented model by modifying the target displacement, 8r. to which the loading
stiffness is directed. This target displacement, which in the basic model is the maximum
displacement of past cycles, 8max' (unless the modifications discussed in Mahin and
Bertero, 1975, apply) can be modified in a manner very similar to that proposed for
strength deterioration. The objective is to increase this target displacement in a manner so
that the reloading stiffness degrades in a cumulative manner expressed by the deterioration
parameter /3. The following model is used to modify the target displacement:
8r,n = (1 + /3) 8r,o = f3k Dt,o

where

(3.4)

8r,n

= new target displacement

8r,o

= old target displacement (previous value of target displacement)

I4

f3k = (1 + {3), with the value of {3 given by Eq. (3.1), using the
appropriate parameters to model accelerated degradation of the
loading stiffness.
The initial value for 81,0 is 8y. and the modification is applied after each inelastic excursion.
When ~Ej approaches Er. the target displacement approaches infinite which causes a
complete loss of stiffness.
The hysteresis response for constant amplitude cycling of a peak oriented model with
accelerated stiffness degradation is shown in Fig. 3.7(a), using E 1 = 100Fy8y (i.e., a=
100) and c = 1.0. When this response is compared to that of the peak oriented model with
strength deterioration, see Fig. 3.7(b), the deterioration patterns appear to be similar.
However, this comparison is misleading as no strength deterioration is associated with the
model shown in Fig. 3.7(a). If in anyone of the repeated cycles the loading would be
continued beyond the displacement of 48y, the reloading stiffness would be maintained
until the original skeleton curve is reached, whereas in the strength deterioration model
shown in Fig. 3. 7 (b) the stiffness would change to the strain hardening stiffness at a
displacement of 48y. The differences between the two models are evident in the two
histories shown in Fig. 3.8, which consist of 4 cycles with amplitude 48y followed by 4
cycles with amplitude 88y.
A computer program has been developed that incorporates all the models identified in
Table 3.1, as well as any desired combination of strength deterioration and stiffness
degradation. The program can trace the hysteretic response for prescribed arbitrary
displacement histories, and is presently being implemented for dynamic time history
analysis. A parameter study is being performed to evaluate the effects of different
deterioration characteristics on the important seismic demand parameters of SDOF systems.
At this time, parameter studies have been performed for bilinear models and peak
oriented models. Referring to Fig. 3.1, these models are described by the yield strength
Fy. the elastic stiffness Ke, and the strain hardening stiffness Kh. The latter is expressed
more conveniently in terms of the strain hardening ratio a= KhfKe. Strain hardening ratios
of 0.0 (elastic-plastic system), 0.02 and 0.1 are used in the parameter study, and all
systems are assumed to have 5% damping.
Time history analyses of these SD 0 F systems were performed to predict the
discussed seismic demand parameters. The natural period of the systems was varied at

15

closely spaced intervals to derive spectral information with a sufficient degree of accuracy
to capture all important trends.
All results discussed in this chapter are based on systems whose yield levels are
adjusted for each record to produce preselected ductility demands of J.l = 1 (elastic
response), 2, 3, 4, 5, 6, and 8. The corresponding yield levels are, therefore, the strength
demands Fy,e and Fy(2) to Fy(8). Thus, all the information generated pertains to systems
that may have different yield levels but equal ductility demands. In the context of the
proposed design approach this is the relevant information, as the available ductility capacity
is the basis for design and the required strength is a quantity derived from the criterion that
the ductility capacity should exceed the ductility demand.
The process of determining the yield level for specified ductilities has to be done very
carefully and may involve several trials because the relationship between yield level and
ductility demand is not necessarily a monotonic function. This is illustrated for one specific
case in Fig. 3.9, taken from Krawinkler and Nassar, 1990-1. For this record and the
selected structural period of 0.9 seconds, there exist at least three yield levels that result in a
ductility demand of 3.0. Clearly, it is only the highest of the three yield levels that defines
the strength demand F y(3). Interpolation between widely spaced F -J.L data points could
have led to very erroneous results. Figure 3.9(b) illustrates that the three systems with
different yield levels have indeed the same ductility demand in the time history response.
The following sections discuss the results obtained from the SDOF study, using three
different sets of ground motion records.

3.3
3.3.1

Statistical Study with Firm Soil Records


Ground Motion Records Used in this Study
The purpose of the study summarized here is to provide demand information that

should prove useful for direct application to design of structures located on rock or firm
soils. Ground motions at such sites depend on seismological/geological parameters such as
earthquake magnitude, stress drop, source-site distance, and travel path through geologic
medium, but are considered to be not greatly affected by local surface geology. Thus, their
frequency characteristics are rather similar in most cases (unless the motions are far from
the source) and, in average, can be represented by a smoothened elastic response spectrum

16

of the shape given in the ATC-3 document for soil types SJ. Thus, 15 records were
selected whose elastic response spectra resemble that of the smoothened ATC-3 ground
motion spectrum. Clearly, there are significant differences in the spectra for each record,
but in average these spectra are a reasonable match of the ATC-3 spectrum, at least in the
range from 0 to 1.0 sec. as will be discussed later.
Basic properties of the 15 records selected for this study are presented in Table 3.2.
The records are from Western U.S. earthquakes ranging in magnitude from 5.7 to 7.7.
The unsealed PGA and PGV values are listed in Table 3.2, together with values obtained
by scaling the PGA to 0.4g and the PGV to 30 em/seconds. The scaling issue is discussed
later. Most of the results presented here are obtained from unsealed records and are
presented in terms of ratios, which makes scaling unnecessary. The strong motion
durations D sm listed in the table are based on the definition given by McCann and Shah,
1979. They define the end of strong motion as the point in time when the time derivative of
the cumulative root mean square function is positive for the last time. The beginning of the
strong motion is obtained similarly by applying the same procedure to the reversed time
history of the record.
Seismic demand parameters for different SDOF systems were computed for each
record and evaluated "statistically". The statistical evaluation is limited to simple arithmetic
mean and mean cr calculations.

3.3.2

Seismic Demands for Bilinear Systems

The basic seismic demand parameters computed for each system are the elastic
strength demand, Fy,e. and the inelastic strength demands, Fy(f.J), which define the required
yield strengths of systems for prescribed ductilities. Since these parameters depend on the
severity of the ground motion, they can be evaluated statistically only once the records are
scaled to a common severity. This will be done later. In nondimensional form, the
inelastic strength demand for constant ductilities can be expressed by the strength reduction
factor Ry(p), which is the ratio of elastic strength demand, Fy,e over inelastic strength
demand, Fy(J.L). This reduction factor can be thought of as an effectiveness factor that
shows by how much the elastic strength demand of a given SDOF system (represented by
the acceleration response spectrum) can be reduced by allowing the system to behave
inelastically within the limits of a predefined ductility ratio.

17

The mean and mean-a values of these R-factors, obtained from the 15 records, for
different ductilities and plotted versus period are shown in Figs. 3.10 and 3.11 for systems
with 0% and 10% strain hardening, respectively. R-J.L-Trelationships of this type, together
with mean elastic response spectra, can be used to evaluate the inelastic strength demands
for SDOF systems on firm soils. (Modifications applicable to MDOF systems are discussed
in Chapter 4). It is evident that these R-J.L-T relationships are highly nonlinear, giving low
R values (i.e., small permissible reductions in yield strength) in the short period range,

peaking in the range of medium periods, and decreasing again for long periods.
These figures show clearly that the often used relationship of R = J.L does not apply
for short period systems but may be a reasonable approximation for medium and long
periods, at least for mean values. It would be easy to show from the data that the often
used relationship R = (2J.L- 1)1/2 is a poor approximation for the short period range. The
data presented here can be used to derive more realistic R-J.L-T relationships, accounting
also for the effect of strain hardening a, which is not negligible ..
The mean-cr diagrams are shown for two reasons. Firstly, a comparison with the
mean diagrams gives an indication of the scatter involved in the data. Secondly, and
perhaps more important, they may be more appropriate for design purposes than the mean
diagrams. The mean- a values of the R-factors correspond to the mean+ a values of the
inelastic strength demands Fy(J.L). Considering the variations in the frequency content of
ground motions, which are responsible for the deviations in strength demands, it may be
prudent to use mean+ crvalues rather than mean values of Fy(J.L). The corresponding mean
-a of the R-factors

are significantly smaller than the means and are smaller than J.L in almost

all cases.
Cumulative damage demands, which affect the ductility capacity, can be represented
in different ways as was discussed in Section 2.3. In most cases they are represented by
energy terms, most appropriately the hysteretic energy dissipation demand, and the number
of inelastic excursions. Statistical data on these parameters are presented in Figs. 3.12 to
3.15.
Normalized hysteretic energy dissipation demands (HE!Fy8y) for different ductilities
and plotted versus period are shown in Figs. 3.12 and 3.13 for bilinear systems with 0%
and 10% strain hardening .. For bilinear systems, normalized hysteretic energy is equal to

L1op/ 8y. where L18pi is the cumulative plastic deformation, a parameter that has been used
for damage modeling by Krawinkler et. al., 1983. If linear damage accumulation is

18

assumed, Fig. 3.13 indicates that for 10% strain hardening systems with equal ductility the
damage in a 0.2 sec. structure is about twice that in a long period structure. If no strain
hardening is present, the differences in normalized hysteretic energy demands between
short and long period structures is much smaller, see Fig. 3.12. Information of this type
can be used to weigh ductility capacities.
As Figs. 3.14 and 3.15 show, for any given ductility the number of inelastic
excursions decreases much more rapi~y with period than the normalized hysteretic energy
demands. Since for bilinear systems the normalized hysteretic energy demand is equal to
(the summation is performed over the number of inelastic excursions), this
observation leads to the conclusion that for equal ductilities the average (mean) normalized
plastic deformation range (I:(Llop/oy)/N) for long period systems must be significantly
larger than for short period systems. This is verified from the results shown in Fig. 3.16.
I:Llopi/Sy,

Some researchers advocate the use of input energy as a basic parameter for assessing
the seismic demand of ground motions. Since for inelastic systems the maximum input
energy occurs usually at the end of the motion (except for short period structures in some
cases) and is then equal to the total dissipated energy (TDE =HE+ DE), it is of interest to
evaluate the contribution of HE to the total dissipated energy TDE. The fraction of TDE
dissipated by HE is shown in Fig. 3.17. From all the parameters investigated in this study
the ratio of HEITDE was the most stable one. The scatter of the data is very small (mean +
a values not shown) and the ratio follows a very consistent pattern, particularly for long
period structures.
In all the results presented here it is evident that the strain hardening ratio a has some

effect on the seismic demand parameters. In almost all cases the inelastic strength demand
for elastic-plastic systems (a = 0) was larger than that of the corresponding strain
hardening system. The reason is that elastic-plastic systems drift more than strain
hardening systems, and a larger yield strength is required in order to limit the ductility to
the same value. The larger Fy(p) for elastic-plastic systems is reflected in smaller reduction
factors. This is documented in Fig. 3.18, which shows the ratio of R-factors of elasticplastic over strain hardening systems. The effect of strain hardening is particularly
pronounced in the hysteretic energy demands for very short period systems, as is illustrated
in Fig. 3.19.
The information presented so far was in normalized form, making scaling of the
records unnecessary. The presentation of actual average strength demand spectra,

19

however, requires scaling of the records before a statistical evaluation can be attempted.
Unfortunately, there is no unique and "best" way to scale records; scaling to common
PGA, or PGV, or equal area under the elastic response spectra has been reported in the
literature, all resulting in different ordinates and shapes of the demand spectra. For the finn
soil records used in this study we have decided to give priority to scaling to a common
PGA of0.4g.

The means and mean+u of the elastic and inelastic strength demand spectra scaled to
this PGA are shown in Figs. 3.20 and 3.21. Superimposed, in dashed lines, is the ATC-3
ground motion spectrum for S 1 soil sites and Av =A a = 0.4. Comparing the elastic
strength demand spectra with this ground motion spectrum, it is noticed that the mean
spectrum matches the ground motion spectrum well in the period range below 1.0 sec.,

~ut

poorly for longer periods. This comes as no surprise as the ATC-3 ground motion
spectrum consists of two bounds, one associated with an effective PGA of 0.4g (the
straight portion) and one associated with a period independent effective PGV of 30 em/sec.
(the curved portion). It is unlikely that a specific ground motion approaches both bounds.
It is probably incidental, but curious, that the mean+uelastic spectrum matches the constant
velocity region of the ATC-3 ground motion spectrum very well.
The important observation to be made from Figs. 3.20 and 3.21 is that the inelastic
strength demand spectra are rather smooth and, for periods larger than about 0.2 sec., can
be represented closely by a smooth curve. Thus, it is quite feasible to develop equations
for inelastic strength demand spectra directly, rather than taking the detour of representing
the elastic spectrum by equations and employing strength reduction factors.
Attempts were also made to scale all records to a common PGV of 12 in./sec. (30
em/sec.). This scaling did not match well with the ATC-3 ground motion spectrum as is
evident from Fig. 3.22.

3.3.3

Seismic Demands for Stiffness Degrading Systems

All the analyses summarized in Section 3.3.2 were repeated for stiffness degrading
system, using the peak oriented (modified Clough) model for stiffness degradation and
keeping all other parameters the same as in the bilinear system. The results were evaluated
for all demand parameters in terms of ratios of degrading to bilinear systems. The most
important result is presented in Fig. 3.23, which shows an example of ratips of strength
reduction factors. Consistently, and for all cases investigated, this ratio was close to 1.0

20

except for very short period systems. The simple conclusion is that for strength design it
makes little or no difference whether the hysteretic behavior of a structure resembles that of
a bilinear system or that of a system with stiffness degradation of the type represented by
the peak oriented model. This is not to say that other types of stiffness degradation (see
Section 3.2) and particularly strength deterioration may not have a significant effect on
strength demands. This issue will be investigated in detail with the models discussed in
Section 3.2.
For cumulative damage parameters, stiffness degradation plays a more important role.
This is illustrated in Fig. 3.24, which shows the ratios for hysteretic energy dissipation
demands. But even here the effect of stiffness degradation is not very large except for
systems with very short periods. One can only come to the conclusion that this type of
stiffness degradation is not a very important issue, except if refined damage models are
used that are very sensistive to the hysteretic energy dissipation.

3.4 Darna2e Potential of Lorna Prieta Ground Motions


[The discussion presented in this section is part of a paper that was submitted for
publication in a special Lorna Prieta issue of the Bulletin of the Seismological Society of
America]
The Lorna Prieta earthquake provided an extensive set of ground motion records and
damage information. In this section the ground motions recorded in this earthquake are
utilized to evaluate basic ground motion and seismic demand parameters, search for
patterns in their characteristics, study the attenuation of the parameters, and use the demand
information together with capacity information on several types of code designed structures
to assess the damage potential of the Lorna Prieta ground motions.

3.4.1

Lorna Prieta Ground Motions


Ground motions in the October 17, 1989 earthquake have been recorded by several

agencies and private organizations, the primary sources being the CDMG and USGS. The
motions recorded by the latter two agencies are well documented (e.g., CDMG, 1989, and
USGS, 1989), many of the records have been digitized and corrected in final format (e.g.,
CDMG, 1990, and USGS, 1990), and several discussions on ground motion issues have
been published already (e.g., Boore et al., 1989, and Competing Against Time, 1990).

21

In the study summarized here, the digitized records made available by CDMG and

USGS were utilized for a global evaluation of strength and ductility demands for bilinear
single degree of freedom (SDOF) systems, with due consideration given to attenuation
characteristics and local site conditions. Records from 51 stations were used for this
purpose. The overriding consideration in the selection of records was that each record
could be viewed as a "free-field" record. Thus, only records from instrument shelters or
one to two story buildings were considered in order to avoid records that could be
considerably contaminated by structural feedback.
The records were classified as either rock, alluvium, or soft site records. This
classification was based in most cases on the information provided in CDMG, 1989, and
USGS, 1990. However, because of the ambiguity of some of the information given in
these references, and in view of the objectives of this study, several judgmental decisions
had to be made in this soil classification. Rock and alluvium site records were used
primarily to study attenuation patterns, and therefore, records which are known to be
uncharacteristic for these two categories were eliminated. This applied to four records that
were classified in CDMG, 1989, as "alluvium" records, but which are known to be soft
soil records (see section on Strength Demands for Ground Motions in Soft Soils) and to the
"rock" motion recorded at the San Francisco Presidio.
Properties of the records classified as rock, alluvium, and other (mostly soft soil site)
records are summarized in Tables 3.3 to 3.5. For, each station only the horizontal
component with the larger PGA value was used. With very few exceptions this component
was also the record with the larger PGV and Arias intensity. The sets of motions for rock
and alluvium sites are rather small for the later described attenuation study, but we had to
limit ourselves to records which were available in digitized form at the time of writing.
The next two sections focus on demand information obtained from time history
response analyses, using these records as input to bilinear SDOF systems with 5%
damping and 10% strain hardening. The emphasis is on strength demand spectra. In the
subsequent section these strength demands are compared with strength capacities of code
designed structures in order to assess the damage potential of the Lorna Prieta ground
motions.

22

3.4.2

Attenuation Characteristics for Rock and Alluvium Sites

Attenuation characteristics are known to follow well established patterns for ground
motions in rock and firm soil (Joyner and Boore, 1988). Thus, the two sets of records
classified as rock and alluvium records were used to evaluate attenuation characteristics of
the peak ground acceleration PGA as well as of the aforementioned strength demands. For
each parameter a regression analysis was performed, using the following relationship
between the attenuating parameter y and the distance r:
logy = a + d log r + kr
where

r0

(3.5)

= (ri + h2)0.5

shortest horizontal distance from station to surface projection of the


rupture surface (km), see Tables 3.3 and 3.4

h
a,d,k

= a depth parameter
= regression parameters.

This equation is of the form proposed by Joyner and Boore, 1988, without
consideration of a site soil correction factor. Joyner and Boore set the value of d to -1.0,
whereas in this study d was used as a free regression parameter. The only constraint on the
regression parameters was that k had to be negative and was taken as zero if it turned out to
be positive in an unconstrained regression analysis. In Joyner and Boore, 1988, the depth
parameter his obtained from a search procedure that minimizes the sum of the squares of
the residuals, and is taken as 8.0 km for attenuation of PGA. This value of 8 km is used
here for attenuation of PGA as well as of strength demand spectra.

Attenuation of PGA
The record sets listed in Tables 3.3 and 3.4 were used in separate least square
regression analyses to evaluate the attenuation of peak ground acceleration PGA. The
regression curves and data points are shown in Fig. 3.25, together with attenuation
relationships obtained by using the Joyner and Boore prediction for a moment magnitude
6.9 earthquake (Boore et al., 1989). As was pointed out already by others (Boore et al.,
1989, and Competing Against Time, 1990), and is confirmed in Fig. 3.25, the
accelerations are higher than predicted by the Joyner and Boore equation and attenuate
slower, particularly at large distances from the fault rupture. A clear difference is also
evident in the attenuation of PGA for rock and alluvium sites, with similar PGA values

23

close to the fault rupture but much slower attenuation for alluvium sites at larger distances,
see Fig. 3.26.

Attenuation of Strength Demand Spectra


Least square regression analysis was performed on period dependent ordinates of the
strength demand spectra, using Eq. (3.5) and the responses computed from the same two
record sets as employed for the attenuation of PGA. Performing the least square fit is
expected to diminish the effects of local site conditions and amplify behavioral patterns in
demand parameters and in the relationships amongst response and ground motion
parameters. The process employed in the regression analysis is simple but very
computation intensive. For each spectral ordinate and for all the variables employed in the
study (e.g., different ductility ratios and periods), the results obtained from the two sets of
records were used for an independent regression analysis. The regression curves for the
spectral ordinates were then used to derive distance dependent spectral information.
Figures 3.27 to 3.31 illustrate representative results obtained from the regression analysis.
'Figure 3.27 shows, in solid lines, the attenuation of the elastic strength demand
(acceleration response) spectra for rock and alluvium sites, using distances to the fault
rupture, r0 , of 10, 20, 40, and 80 km. Also shown, in dashed lines, are the 1988 UBC
code design values ZC for soil type S1 in Fig. 3.27(a) (rock) and S2 in Fig. 3.27(b)
(alluvium). Comparing these curves with the elastic spectra gives an indication how close
the elastic strength demand imposed by the Lorna Prieta ground motions is to the code
"design value."
There are clear differences in the shapes and attenuation characteristics of the elastic
spectra for rock and alluvium sites. The near-source (10 km) rock spectrum exhibits one
large peak at 0.3 sec. and rapid decay at longer periods, whereas the corresponding
alluvium spectrum exhibits a shon period plateau and slower decay at longer periods. For
both the rock and alluvium spectra the attenuation with distance is rapid in the shon period
range whereas it is very slow for specific longer periods. This phenomenon results in a
significant change in spectral shapes as a function of distance, as is illustrated in Fig. 3.28,
which shows the same spectra but normalized with respect to the regressed PGA at the
specified distances. Peaks in the spectra diminish with distance and are replaced by wider
plateaus, resulting in almost constant spectral values from 0.2 to 1.0 second periods at a
distance of 80 km (San Francisco). These phenomena are more pronounced for rock sites
than alluvium sites.

24

The significant changes in shapes of the rock spectra with distance from the fault
rupture may not be critical for the design of structures located on rock sites, since the
design likely will be governed by spectra derived from near-site earthquakes. However,
there is a critical consequence to be considered. The elastic spectra shown in Figs. 3.27
and 3.28 represent frequency characteristics of the ground motions, and a broad plateau
with relatively high spectral values at longer periods clearly indicates that soft soils over
rock will be much more excited than anticipated from a typical near-source rock spectrum
with peaks at short periods. There is no doubt that the slow attenuation of long period
components of the ground motion, reflected in the wide plateaus of the far-source rock
spectra (80 km), has contributed significantly to the amplification of motion in soft soils in
and around San Francisco and Oakland. This conclusion is not new, but little
documentation existed so far on the importance of this effect. As Fig. 3.28(a) shows, the
spectral accelerations at long periods in far-source rock spectra may be easily three times as
large as anticipated from near-source rock spectra.
Figures 3.29 and 3.30 show results similar to those of Figs. 3.27 and 3.28, but for
inelastic strength demand spectra for Jl = 4. As was stated previously, these spectra
identify the required strength of SDOF systems whose ductility is limited to a value of 4.0.
The shapes of these spectra are relatively smooth and are very different from the elastic
ones, indicating that the inelastic strength demands are not related to the elastic ones by a
period independent reduction factor (R-factor) as is assumed in present design codes. The
variation in spectra shapes with distance is illustrated in Fig. 3.30 and shows only minor
variations for alluvium spectra, but great variations for rock spectra based on the much
slower attenuation of inelastic strength demands for longer periods.
Figure 3.31 shows regressed strength demand spectra for Jl = 1 (elastic), 2, 3, and 4
at a distance to the fault rupture of 10 km. Again, the nonlinear relationship between elastic
and inelastic spectra is evident, particularly in the rock spectrum where the peak in the
elastic spectrum disappears already in the Jl = 2 spectrum.

Local Variation of Ground Motions


Much has been said and written about the great variation in ground motions
throughout the Bay Area. However, the following results obtained from an evaluation of
localized damage potential indicate that seismic demands derived from closely spaced
ground motions are very similar unless great variations in site conditions exist. At or near
Stanford University, records were obtained from the four stations listed in Table 3.6. Only

25

one of the records (SLAC) can be viewed as a free-field ground motion, the other three
motions were recorded in the basement or ground level of multi-story buildings. The
SLAC station is located on tertiary sedimentary rock, whereas the other three stations are
on alluvial soils.
The elastic strength demand spectra for the four records are shown in Fig. 3.32.
With all the differences in peak ground motion values and site conditions it comes as some
surprise that the elastic strength demand (acceleration response) spectra are so similar over
the whole range of periods. One may want to conclude that the engineering content of the
ground motions is similar and the seismic demands on the Stanford campus and its vicinity
are rather uniform.

3.4.3

Strength Demands for Ground Motions on Soft Soils

It is well established by now that records at soft soil sites (particularly fill on bay
mud) show great amplifications of PGA values compared to nearby firm soil records.
However, it would be misleading to proportion seismic demands imposed on structures
according to PGA alone, without regard to frequency characteristics of the ground motions.
The following discussion focuses on an evaluation of the damage potential of Lorna Prieta
ground motions that were recorded on soft soil, utilizing the aforementioned elastic and
inelastic strength demand spectra. Results from six individual ground motions (see Table
3.7) are discussed and illustrated, since the records are site specific and no statistical
evaluation is possible. Five of the six ground motions were recorded on deep soft soils
whereas .the Stanford/SLAC record is representative of motion in firm ground. CDMG,
1989, classifies four of the five soft soil records under "alluvium," but all these stations are
on soft soils and, in addition, at one site (Oakland) several feet of fill material were placed
on top of fine-grained sand.
Elastic as well as inelastic strength demand spectra for each of the records are shown
in Fig. 3.33. Superimposed on the spectra are two sets of curves that relate the spectra to
code design considerations. The dashed curves (S2 - S4 ) represent the 1988 UBC code
design coefficients ZC for soil types S 2 , S3 , and S4 , and are a measure of the elastic
strength demand envisioned by the code. The dashed-dotted curves (SMF88 and CMF68)
will be discussed in the next section. An inspection of this figure as well as of Figs. 3.34
to 3.36 lead to the following conclusions on the seismic demands imposed by ground
motions on soft soil.

26

1. The shapes of the elastic strength demand (acceleration) spectra, which are shown in
Fig. 3.34 (spectra normalized with respect to PGA, or dynamic amplification factors,
DAF) are very different for all six records. Except for the Stanford/SLAC motion,
which was recorded on firm ground, the elastic spectra show a clear signature of the
soft and deep soil on which the motions were recorded. This signature is most evident
in the large DAF of the Foster City I Redwood Shores record at a period of 0.7 seconds
(DAF = 4.0) and in the wide humps of several of the other spectra, extending to and

beyond a period of 1.5 seconds in the Emeryville and Oakland spectra


2. All six records, including the Stanford/SLAC record, exhibit a low high-frequency
content, represented by the small DAFs at periods shorter than 0.25 seconds. This may
help to explain the relatively "good" performance of most short-period structures, such
as low-rise unreinforced masonry and tilt-up structures, in the vicinity of the record
locations.
3. As Fig. 3.33 shows, with few exceptions the elastic strength demand imposed by the
Lorna Prieta earthquake was smaller, and in many cases much smaller, than the strength
demand implied by the 1988 UBC for a design earthquake (probably soil type S4
should be used for all records except Stanford/SLAC). It must be concluded that in this
earthquake most structures did not experience the severity of motion for which modem
codes intend to provide protection against collapse, and we should derive no comfort
from the observation that modem structures survived this earthquake without collapse.
4. For many soft soil sites it must be concluded that the code prescribed S4 elastic force
demand is too low in the period ranges in which soil response magnifies the DAF. For
instance, if the elastic Emeryville spectrum is scaled up to a PGA of 0.4g, the spectrum
will exceed the S4 curve considerably in the period range of 1.0 to 1.5 seconds. Site
specific design spectra need to be utilized for such soil sites.
5. The shapes of the inelastic strength demand spectra for J1. = 2, 3, and 4 are quite
different from those of the elastic strength demand spectra. The ratio of elastic to
inelastic strength demand (the strength reduction factor R) is by no means constant; it is
much smaller than J1. for short periods and for periods of low elastic strength demands
preceding ranges of high elastic strength demands (humps in the elastic spectra). The
reason for the latter is that the effective period of the inelastic system elongates and
shifts into period ranges of high strength demands. This is evident from Fig. 3.35,
which shows plots of the strength reduction factor for the Foster City I Redwood

27

Shores record for p. =2, 3, and 4. The peaks and valleys of these plots coincide with
those of the corresponding elastic strength demand spectrum. The consequence of this
phenomenon is that the peaks evident in the elastic spectra diminish and ultimately
disappear in inelastic strength demand spectra for increasing p. values.
6. The observation made in the previous paragraph has significant implications for design.
For structures with small ductility capacity the required strength will be high and the
elastic strength demand spectrum will be an important design parameter. For structures
with large ductility capacity the inelastic strength demand spectra, which are very
different from the elastic spectra, should control the design. This implies that more
knowledge. needs to be acquired on the magnitude and shape of inelastic strength
demand spectra. The data discussed here show that the shape is site dependent, but in a
different manner than the elastic spectra. This can be seen by comparing the normalized
elastic strength demand spectra of Fig. 3.34 with the normalized inelastic strength
demand spectra for p. = 4 shown in Fig. 3.36.
7. The elastic strength demand imposed on structures on soft soil may be easily 6 times as
high as that on nearby structures built on rock. This factor is obtained as the product of
PGA ratios and DAF ratios of soft-site to rock-site records.
The work discussed here led to an interesting observation worthy of reporting. The
six records used in this section were initially hand-digitized four days after the Lorna Prieta
earthquake from enlarged photocopies of the analog traces published by USGS and
CDMG. No filtering and baseline corrections were applied to the hand-digitized records
and results similar to those shown in Fig. 3.33 were available five days after the
earthquake. It is of some interest to note that the spectra derived from these cursory
digitized records are not much different from those presented in Fig. 3.33. An illustration
of typical differences is shown in Fig. 3.37. For reasons that would deserve explanations
the differences are larger for the inelastic spectra than for the elastic ones, and are larger in
the long period domain than in the short period one.

3.4.4

Damage Potential of Ground Motions


In order to assess the damage potential of the Lorna Prieta ground motions, the

strength capacities of structures must be compared to the strength demands imposed by the
ground motions. These capacities must be estimated with due consideration given to the

28

real strength of structures. For many reasons structures are stronger, and sometimes much
stronger, than is implied by the seismic base shear coefficients used in code design. The
overstrength above the design force level may come from many sources, and depends
primarily on the type of structural system and the number of stories (period) of the
building. The strength of five types of code designed buildings, considering most but not
all sources of overstrength, was estimated by Osteraas and Krawinkler, 1990, and Nassar
and Krawinkler, 1991, and is illustrated in Fig. 3.38. Four of these five types of
structures were designed in conformance with the 1988 SEAOC Bluebook (SBF88 = steel
braced frames, SMF88 = steel moment frames, SPF88 = steel perimeter frames, and
CMF88 =concrete moment frames), and one type was designed in conformance with the
1968 SEAOC Bluebook (CMF68 =concrete moment frames).
The strength capacity curves of the strongest and weakest structure types (SMF88
and CMF68, respectively) provide a range of the expected strength of modern code
designed structures and are used here to assess the damage potential of the Lorna Prieta
ground motions. These two capacity curves are superimposed on the strength demand
spectra illustrated in Figs. 3.27, 3.29, 3.31, and 3.33, and are left to the reader for
interpretation, with only a few observations summarized here.
Figure 3.27 indicates that even for code designed structures located on rock, small
inelastic deformation demands have to be expected in some cases as far away as 80 km
from the fault rupture, considering that the CMF68 curve falls partly below the elastic
strength demand spectra for this distance. Figure 3.29 indicates that global ductility
demands of 4 are anticipated as far away as 20 km from the fault rupture. Figure 3.31
indicates that at a distance from the fault rupture of 10 km the ductility demands could have
exceeded the value of 4 considerably for CMF68 structures with period between 0.2 and
1.0 seconds and located on alluvial soils. Clearly, these observations are more qualitative
than quantitative, since the spectra are obtained from a regression analysis and represent
average rather than site specific conditions, and since significant deviations from the
estimated values of structure strength have to be anticipated.
Figure 3.33 shows clearly that in soft soils, where the shaking was severe even at
large distances, the ductility demands for modern structures could have been substantial.
This holds true particularly for early-70 vintage low- and mid-rise reinforced concrete
frame buildings (CMF68), except for 2-story buildings. The global ductility demand for
modern steel moment frames (SMF88) was small for low-rise and high-rise construction
and moderate for mid-rise buildings.

29

The demand and capacity curves shown in Figs. 3.27, 3.29, 3.31, and 3.33 illustrate
the variation of expected damage (assuming that ductility is an acceptable measure of
damage) with source-site distance and site soil conditions - from the viewpoint of demands,
and types of structural system and number of stories (period) - from the viewpoint of
capacities. In none of the cases illustrated, the ductility demand is severe enough to justify
collapse of a modern structure. But relatively high ductility demands are evident in some
cases even though the ground motions were significantly smaller (except close to the
source) than anticipated by present code philosophy. This indicates that excessive ductility
demands may have to be anticipated in some code designed structures if the ground
motions approach the level envisioned in design codes. This holds true particularly for
structures located on soft soils. It is understood that soft soil PGA amplification decreases
as the rock PGA increases (Idriss, 1990, estimates that at rock PGAs exceeding 0.4g no
soft soil PGA amplification occurs), but there is good reason to believe that significantly
larger soft soil strength demands can be generated in a severe earthquake than those shown
in Fig. 3.33 and, therefore, the ductility demands for modern structures may be much
higher than those shown in this figure.
In summary, the Lorna Prieta earthquake has again demonstrated the sensitivity of
ground motions to source-site distance and orientation, travel path through geologic media,
and local site conditions. Moreover, the earthquake has also demonstrated the great
dependence of the elastic as well as inelastic structural response on the frequency
characteristics of the ground motion. The simple conclusion is that the demands imposed
by an earthquake on a structure, which should be the basis for protective design, must be
evaluated with due consideration given to all the aforementioned factors. This conclusion
points towards the need for detailed microzonation and raises questions on the
appropriateness of presently employed global seismic zoning. It is recognized that detailed
microzonation is a long way from reality because much of the needed information is not yet
available. However, the Lorna Prieta earthquake provided an excellent opportunity to
assess seismic demands based on recorded ground motions, as a first step in identifying the
sensitivity of the demands to known site conditions and structural response characteristics.

30

3.5 Seismic Demands for Selected Records from Japan


Eight Japanese ground motion records were provided by the Kajima researchers for
demand evaluation. Basic data on these records are summarized in Table 3.8. The
unsealed records are of very different severities and frequency contents, and cannot be used
for a statistical evaluation. F9r purposes of direct comparison all records were scaled to a
common peak velocity of 50 em/sec. The PGA values of the scaled records are also
summarized in Table 3.8. Records scaled to a PGV of 50 em/sec. are often employed in
Japanese practice for Level 2 design (catastrophic earthquake) in highly seismic regions
such as Tokyo.
These 8 records were used as input to the same SDOF systems used in the study
discussed in Section 3.3. Some of the results for bilinear systems with 5% damping and
10% strain hardening are illustrated in Figs. 3.39 to 3.42. Again, it must be emphasized
that the results are always presented for specified ductility ratios and, therefore, the
strengths of the systems vary accordingly. All presented results are for records scaled to a
PGV of 50 em/sec., which may help to assess the importance of the individual records for
Level 2 design.
For two of the scaled records (Tokyo 101 and Hachinohe NS) the strength demand
spectra for J.l = 1 to 8 are shown in Fig. 3.39. The demands are very different for the two
records, the Tokyo 101 record controlling elastic and inelastic strength demands for short
period systems and the Hachinohe NS record controlling for most of the long period range.
The relevance of each record for strength design is more evident from Fig. 3.40,
which shows the elastic strength demands as well as the inelastic strength demands for J.l =
2, 3, and 4 for all 8 records. Depending on the degree of tolerated inelastic behavior (target
ductility), different records control in different period ranges. For elastic systems
(probably the least relevant case for Level 2 design), the required strength is controlled by
the Osaka 205 record for short period systems, the Sendai 030 and Nagoya 306 records for
medium period systems, and the Hachinohe NS record for long period systems (see Fig.
3.40(a)). All other records do not control over a significant period range. [As a matter of
code comparison, if one places a rough skeleton curve to all elastic spectra, the resulting
curve is above the ATC-3 spectrum for soil type S3 ].
The importance of the records changes significantly with the degree of inelastic
behavior, as can be seen from Figs. 3.40(b) to (d). The Tokyo 101 record becomes the

31

only important one for systems of less than 0.8 sees., the Sendai 030 records maintains
some relevance for systems of medium periods, and the Nagoya 306 record takes on
increasing importance for medium to long period systems. The range of importance of the
Hachinohe NS record is pushed to very long period systems. Again, all other records do
not control over a significant period range. Even though these observations are based on
SDOF strength demand spectra, which pay no attention to multi-mode effects, there is little

doubt that some of the other records can be eliminated from considerations- for strength
design.
A different picture of relative importance of the records is obtained when cumulative
damage parameters are evaluated and the duration of ground motion becomes an important
consideration. Let us again assume that normalized hysteretic energy dissipation demand
(HE!Fyay) is the relevant parameter for this purpose. Plots of this parameter for the 8
records and J.L = 2 and 4 are shown in Fig. 3.41. Different records gain now on
importance, particularly the two Hachinohe records (NS and EW), which are characterized
by long strong motion durations and significant frequency content at long periods. On the
other hand, the Tokyo 101 record, which is of short duration, causes very little cumulative
damage even for short period systems where it controlled the strength demand.
In the interpretation of these results it must be considered that Fig. 3.41 shows
hysteretic energy dissipation normalized by Fy8y A very different picture is obtained for
the absolute values of energy dissipation, as is illustrated in Fig. 3.42. This figure shows,
for the same cases as illustrated in Fig. 3.41, the hysteretic energy dissipation demands per
unit mass. Only two records control this demand, the Sendai 030 motion for periods less
than 1. 7 seconds, and the Hachinohe NS motion for larger periods. The differences
between Figs. 3.41 and 3.42 come from the fact that in these respective period ranges these
two records demand high strength as well as many inelastic cycles if the ductility is to be
limited to the specified values for which the plots apply.
The seismic demand assessment for the 8 Japanese records discussed in this section
is intended to put the importance of these records in perspective, in the context in which
these records are believed to be used in Japan. The records have vastly different frequency
contents, which in part are caused by greatly different site soil effects, and are very unlikely
to occur at the same site. As far as frequency content is concerned, they appear to bracket a
wide range of different conditions. The selection of specific records for particular designs
will depend on how much is known about the local conditions. The presented information
is intended to assist in assessing the importance of the individual records for use in design.

32

3.6

A Pilot Study on Cumulative Damage

[This section was prepared by Kajima researchers]

Experiences from past strong earthquakes have shown the vulnerability of reinforced
concrete buildings to strong ground motions. In order to assess the seismic safety of
reinforced concrete buildings, quantitative analysis of structural damage under earthquakes
is essential. Several damage parameters for reinforced concrete buildings have been
proposed for this purpose. The objective of this study is to evaluate the damage potential
of recorded earthquake ground motions by analyzing one of the damage parameters of
single degree of freedom systems.
Damage Index
Under earthquake loading, reinforced concrete structures are generally damaged by
both repeated stress reversals and high stress excursions. The damage index proposed by
Park and Ang is expressed as a linear combination of the maximum deformation and the
hysteretic energy:

in which

damage index, an empirical measure of damage


(D > 1 indicates total damage or collapse)

Om

the maximum response deformation

Ou

ultimate deformation capacity under static loading

Qy

calculated yield strength

dE

=
=

incremental dissipated hysteretic energy

coefficient for cyclic loading effect

The limiting value of D (i.e., damage capacity of a member) is lognormally


distributed with a mean of approximately 1.0 and a standard deviation of 0.54 (Park and
An g).

33

Recorded Earthquake Ground Motions


The following recorded earthquake ground motions are used in this study:
a)
b)
c)

El Centro
Hachinohe
Taft

The acceleration time histories, Fourier amplitude spectra, and linear elastic response
spectra of these earthquakes are shown in Figures 3.44 to 3.46.
SDOF Models (see Fig. 3.43)
The following SDOF models are utilized in the analysis:
a)
b)

basic bilinear model


peak oriented degrading type bilinear model

Input Parameters for Analysis

kp
h
8y

SDOF system

~
Ou

damage index

Qy
8y:
Oemax:

O.lk0

0.05
0.58emax
0.05
4.08y
k0 8y

the yield deformation


the maximum elastic deformation

Results
Damage indexes for these SDOF systems for El Centro, Hachinohe, and Taft are
shown in Figure 3.47 to 3.49, respectively. Though the yield deformation is assumed as
half of the maximum elastic deformation of the SDOF system (i.e., each SDOF system has
a strength reduction factor of 2.0), the damage indexes are different. This shows the
difference of the damage potential among earthquakes. For example, among these
earthquakes, Hachinohe has comparatively high damage potential around 1.5 second of
natural period.

34

4. SEISMIC DEMANDS FOR MULTI DEGREE OF FREEDOM SYSTEMS

The previous chapter provided information on seismic demands for inelastic SDOF
systems. This information is relevant for relative assessment of damage potential of
ground motions, but needs to be modified to become of direct use for design of real
structures, which are mostly multi-degree-of-freedom (MDOF) systems governed by
several translational and torsional modes. For elastic MDOF systems, the combination of
modal responses using SRSS, CQC, or other approaches, provides reasonable estimates of
peak dynamic response characteristics. For inelastic MDOF systems, modal superposition
cannot be applied with any degree of confidence and different techniques have to be
employed in order to predict strength or ductility demands that can be used for design.
Story ductility ratio, defined as the maximum dynamic interstory displacement
normalized by the interstory yield displacement, is used here as the basic deformation
parameter and, therefore, the strength demand associated with the a specific story ductility
ratio becomes the basic design parameter. This strength demand differs from that of the
corresponding SDOF system because of the higher mode and torsional effects, and
depends on many other structural characteristics such as the member stiffness and strength
distribution over the height of the structure, structural system redundancy and modes of
failure of critical structural elements.

4.1 MDOF Models Used in this Study


The research discussed in this chapter is intended to provide some of the answers
needed to assess strength demands for inelastic MDOF systems for comparison with their
SDOF counterparts. The focus is on a statistical evaluation of systems that are regular from
the perspective of elastic dynamic behavior. Realizing that no two structures are the same,
and that the dynamic behavior of real structures depend on so many parameters, simplified
MDOF models were used in this study to gain insight into their inelastic dynamic behavior.
Therefore, for convenience in computer analysis, "regular" 2-dimensional single-bay

frames with widely spaced elastic modal periods were utilized to represent the behavior of
three distinct structural models. The torsional effects of 3-dimensional structures are
neglected and the results do not apply to structures with closely spaced modal periods.
The three models used are illustrated in Fig. 4.1. The "beam hinge" (strong columnweak beam) model, referred to as BH model from here on, represents structures (e.g.,

35

special moment resisting frames) in which plastic hinges will form in beams only (as well
as supports). The "column hinge" (weak column- strong beam) model, referred to as CH
model from here on, represents structures (e.g., braced frames and moment resisting
frames designed to accommodate plastic hinges in columns) in which plastic hinges will
form in columns only. The "weak story" model, referred to as WS model from here on,
represents structures in which plastic hinges will form in columns of the first story only.
This model represents the behavior of frames with a strength (not stiffness) discontinuity in
the first story. Note that structures with infill walls in all but the first story will have both
strength and stiffness discontinuities, which is not represented by the WS model.
The three models were designed to develop the structure mechanisms shown in Fig.
4.1 under the 1988 UBC equivalent static lateral load pattern, i.e., relative member
strengths were tuned so that all shown plastic hinges develop simultaneously under this
specific lateral load pattern. Note that the BH model cannot develop a story mechanism,
the CH can develop one in any story and theWS can develop one only in the first story.
The only difference between the CH and WS models is the strength discontinuity above the
first story for the WS model, in which all stories remain elastic throughout the entire
dynamic time history analysis discussed later.
The elastic member stiffnesses in each story are tuned so that, under the 1988 UBC
equivalent static load pattern, the interstory drift in every story is identical, resulting in a
straight line deflected shape. Furthermore, the stiffnesses are tuned such that the first mode
period of each structure is equal to that given by the UBC as T = 0.02hn3!4 sec, where
hn = total frame height in feet. As a consequence, the first mode shape for the three models

is close to a straight line. This tuning is done for the three models in order to permit a
direct comparison of dynamic analysis results. Only structures with 2, 5, 10, 20, 30 and
40 stories (story height h = 12'), ranging in period from T = 0.217 to 2.051 sec, were
studied. Pertinent data for the first five elastic modes of the CH (and WS) models are
presented in Table 4.1. The modal periods for the BH models are the same for the first
mode but deviate slightly for higher modes due to difficulties in precise stiffness tuning,
which also led to slight deviation in modal masses.
A bilinear moment-rotation (M-8) hysteresis model with strain hardening ratios a= 0,
2 and 10% is assumed at each plastic hinge location. Therefore, under an incremental
("push-over") 1988 UBC load pattern, each structure will have a bilinear response identical
to that of a bilinear SDOF system, which was employed as one of the SDOF models in
Chapter 3.

36

The base shear strength, Vy, is varied for each structure and ground motion record so
that it is identical to the inelastic strength demand Fy(J.Lt) of the corresponding SDOF system
with the first mode period and the same strain hardening ratio, for SDOF target ductility
ratios of J..Lt = 1, 2, 3, 4, 5, 6 and 8. Tuning the base shear capacity of MDOF systems to
the strength of their corresponding SDOF systems subjected to the same ground motion
record, is critical in order to compare results for SDOF and MDOF systems. Because
modal superposition is hardly feasible for inelastic MDOF systems, it is desirable to utilize

SDOF demand estimates to predict MDOF demands. The inelastic strength demands can be
evaluated for SDOF systems as discussed in Chapter 3. The question to be answered is
how different are the ductility demands fo:r MDOF systems if their base shear strengths are
identical to the inelastic strength demands of the SDOF systems for the prescribed target
ductility ratio f.Lt Or more relevant for design, the question becomes: how different should
the base shear strength of the MDOF system be (assuming the code prescribed seismic load
pattern over the height of the structure), compared to that of the corresponding SDOF
system, in order to limit the maximum story ductility ratio in the MDOF system to the
specified target ductility ratio J..Lt (see Fig. 4.2).
The following assumptions were made in the design and analysis of the MDOF models :

The mass of each story is the same.

The height of each story is the same.

Concentrated plasticity at the ends of members, i.e., no yielding occurs within


members.

Even though the 1988 UBC uses different equations for estimating the natural
periods of different structures, the same equation, T

= 0.02hn3/4 sec is used for

the three models for comparison.

The P-delta effect of gravity loads is not included explicitly and is assumed to be
accounted for through a reduction in the strain hardening ratio a.

The effect of gravity load moments on plastic hinge formation is not included.
This simplifies the nonlinear structural response from a multi-linear to a bilinear
one.

37

Axial load - bending moment (P-M) interaction is not considered, i.e., the
bending strength at each plastic hinge location is assumed to be constant,
irrespective of the axial load carried by a column.

The overturning moment at a given story is assumed to be the integral area of the
story shears from the top of the structure to the base of the columns of that story
at a given instance. No allowance was made for moments at the base of columns.

Therefore, the calculated overturning moments are the externally applied moments
irrespective of the type of model used.

5% damping was assumed for the frrst two modes in the time history analysis.

The BH and CH models were designed and utilized in a previous study to assess
MDOF strength and ductility demands for the 1987 Whittier Narrows ground motions
(Krawinkler and Nassar, 1990). In the study summarized here these two models as well as
the WS model are utilized to derive statistical (mean and mean+cr) values for MDOF
strength and ductility demands for S1 type ground motions, utilizing the 15 S1 ground
motions for which SDOF results are available. All model structures, with their strength
tuned as discussed previously, were subjected to the 15 S1 motions using a modified
version of the DRAIN-2D analysis program. The results of the analyses were evaluated for
strength and ductility demands as well as for overturning moment characteristics.
A total of 5,670 nonlinear time history analyses were performed, using 15 ground
motion records, 3 types of structures (BH, CH, and WS), 6 different fundamental periods
(number of stories = 2, 5, 10, 20, 30, and 40), 7 yield levels (corresponding to SDOF
yield strength for target ductility ratios J..L1 = 1, 2, 3, 4, 5, 6, and 8), and 3 strain hardening
ratios (a= 0.0, 0.02, and 0.10). In addition, elastic time history analyses (J..L =e) were
performed as needed. Note that an MDOF system designed for a base shear equal to that
imposed on an equivalent elastic SDOF will not necessarily remain elastic, therefore the
distinction between J..Lt = 1 (elastic SDOF) and J..L = e (elastic MDOF). From each dynamic
time history analysis the following information was stored for evaluation :

Story ductility ratios, J.li

Maximum shear force in each story, Vi

Maximum overturning moment in each story, MoT,i


The maximum overturning moment in a story is not necessarily the same as the

overturning moments resulting from maximum story shears since the story shear maxima

38

may not occur at the same instance. Figures. in which the distribution of the
aforementioned parameters are shown over the height of the structure. represent envelopes
of story maxima for these parameters. irrespective of the time at which they happen.
Results presented in the following sections are mean values obtained from the 15 S1
records, a record set designated from here on as 15s.

4.2 Ductility and Base Shear Demands


If an MDOF system were to behave exactly like its equivalent SDOF system, all

resulting interstory ductility ratios would be equal to the SDOF target ductility ratio Jlt
However, due to higher mode participation in MDOF systems, this is usually not the case.
The question is how do the MDOF interstory ductility ratios compare to those of the
equivalent SDOF system. Or more important, how much do we need to modify the base
shear capacity of the MDOF system, compared to that of the equivalent SDOF system, to
limit the maximum interstory ductility ratio to the same prescribed value (see Fig. 4.2).

4.2.1

Story Ductility Demands


Fig. 4.3 (a) shows an example of the mean a story ductility demand distribution for

20-story BH structures, for an SDOF target ductility ratio of J.lt = 8 and strain hardening a
= 0% for the 15s data set (thus the notation 15s.bh20-8.00). Part (b) of the figure shows
the corresponding information for the CH model.
Fig. 4.4 shows the mean story ductility ratios for the three models (BH, CHand WS,
respectively) with 0% strain hardening, base shear capacities equal to the inelastic strength
demand for SDOF systems with a target ductility ratio Jlt

= 2.

Fig. 4.5 is similar to Fig.

4.4 except that the target ductility ratio J.lt is equal to 8. Fig. 4.6 is similar to Fig. 4.5
except for a= 10%.
If the MDOF structures would respond like SDOF systems, the story ductility ratio in

every story would be equal to the target ductility ratio Jlt Clearly, this is not the case, as
expected, because of higher mode participation and differences in mechanisms in the three
structure types. The results in Figs. 4.4 to 4.6 show consistent trends that can be
summarized as follows :

39

The pattern of the story ductility demand distribution over the height of the structure is
similar for the BH and CH structures. The distribution is rather uniform over height
for 2 to 10 story structures; for taller structures the ductility demands are high in the
lower and upper stories (except for very large ductility ratios (p.1 = 8)), and decrease
significantly in the middle stories. This pattern is due to the participation of higher
modes whose elastic periods (see Table 4.1) coincide with the peaks of the elastic and
inelastic strength demand spectra.

The ductility demands for the CH structures are consistently higher than the demands
for the BH structures, particularly in the lower stories, even though both structures are
designed for the same base shear strength. This is because a story mechanism can
readily develop in a CH model, resulting in more drift.

The deviation of story ductility ratios from those predicted by the equivalent SDOF
systems increases with the number of stories, the target ductility ratio, and decreasing
strain hardening. The largest deviation is associated with the WS model and the least
with the BH model.

The strain hardening ratio a has a significant effect on the ductility demands, as can be
seen by comparing the results for a = 0 and 10%. The elastic-perfectly plastic
structures (a= 0%) create much larger ductility demands, particularly in the lower
stories of tall structures, even though the base shear strength, Vy
is greater than that for a = 10%.

= Fy(Jl.r), for a= 0%

The relative story strengths, which are in conformance with the 1988 UBC seismic load
pattern, do not cause an equal ductility demand in each story. (It probably was not
intended to do so).

In most cases, the maximum story ductility ratio occurs in the first story, indicating that

the 1988 UBC seismic load pattern generally provides good protection against
excessive ductility demands in upper stories. There are notable exceptions, particularly
for 5 to 20 story structures and lower ductility ratios, where the story ductility demands
in the upper stories are sometimes higher than in the first story. This indicates the need
for some (but not major) revisions of the code seismic load pattern. In the evaluation of
the results discussed later it is assumed that the first story ductility ratio governs for
design and that the required base shear strength is determined from this first story
ductility ratio.

40

The ductility demands for theWS model demonstrate the great problems inherent in
these structures. Only the first story ductility ratios are actual ductility demandsl, and
they are much larger than those of the corresponding CH structures. It should be
emphasized that in the first story, the WS and CH structures are identical, i.e., they
have the same strength and story mechanism. Differences start above the first story,
insofar that all other stories in the WS structures are "strong" and remain elastic
whereas the strength of their counterparts in the CH structures are tuned to the first
story strength and the code seismic load pattern. The high ductility demand imposed on
the WS structures is attributed to the concentration of energy dissipation in the first
story. The demand is even higher for higher ductility ratios and no strain hardening.

Even though the shear transfer in the WS structure is limited by the shear capacity of
the first story, the elastic vibration of the upper stories can attract severe shear forces
(not limited by any yield level) that transmit high axial forces to the first story. It is
worth noting that the stories above the first do not vibrate as a rigid body, and that their
interstory drift demand increases with strain hardening in the first story.

Strain hardening helps distribute the seismic energy throughout the entire height of the
BH and CH models. Comparing Figs. 4.5 and 4.6 shows that the double curvature
noted in the story ductility demand distribution of tall structures with no strain
hardening, almost disappears. The upward transfer of shear, and consequently seismic
energy, increases with strain hardening.

Figure 4.7 shows the story ductility demands in the first story, J.LJ, of the MDOF
systems for SDOF target ductility ratios J.Lt = 2 and 8 and strain hardening ratios a= 0 and
10%, plotted against the first mode period of the MDOF systems. The target ductility
ratios, indicated by the horizontal dashed lines, are the ductility ratios of the SDOF systems
whose strength, Fy(J.Lc), was used as base shear strength Vy in the design of the MDOF
systems. The following can be noted from Fig. 4.7:
1 "Ductility" values larger than 1.0 are shown in the upper stories of the

WS structures
to illustrate the increase in elastic forces, attracted in these stories, beyond the force
level based on the code seismic load pattern. For instance, as shown in Fig. 5.5 (c), the
elastic shear forces in the upper stories of the 40~story WS structure (with target
ductility J.Lt= 8 and a= 0%), is about 6 times that required by the code load pattern,
even though the shear transfer to the upper stories is limited by the strength capacity of
the first story. This increase is mainly caused by increased elastic vibrations above the
first story.

41

The deviations from the target ductility ratios exhibit consistent trends, being small for
low-rise BH and CH structures and increase with the number of stories. However, the

WS structures exhibit very large ductility demands. Also, the difference in ductility
demands for CHand BH models appears to be period independent (almost parallel lines
in the figure). The maximum ductility demand for the 40 story structures, may be as
high as 10 times the target ductility ratio for theWS model, 3 times for the CH model
and 2 times for the BH model.

The significant effect of strain hardening is evident by comparing the results for strain
hardening a= 0 and 10%. Systems with strain hardening (a= 10%) give consistently
smaller ductility demands than those without (a= 0%). This is attributed to the fact
that systems without strain hardening drift more and especially so for systems that can
develop separate story mechanisms, i.e., WS and CH models.

4. 2. 2 Base Shear Demands


The results illustrated so far, clearly show that the ductility demands for MDOF
systems differ significantly from those of the corresponding SDOF systems. The
implication is that the base shear strength of the MDOF systems must be modified
compared to that of the SDOF systems in order to limit the story ductility demand to the
desired target values (see Fig. 4.2). The following discussion provides an approach to
achieve this objective and illustrates several results. It needs to be emphasized that the
results presented for the 15s record set are based on data derived from 15 S1 ground
motions, therefore, they provide representative information for design of MDOF systems
located on firm soils or rock. Also, results have to be put in context of the aforementioned
assumptions listed Section 4.1.
For SDOF systems, the parameter that relates elastic to inelastic strength demands is
the strength reduction factor Ry(J.Lr) = Fy.e/Fy(J.Lr). It is a convenient design parameter as it
permits, in concept, the derivation of a strength design spectrum from an acceleration
response (elastic strength demand) spectrum. The advantage of using strength reduction
factors, which are strength demand ratios, is that they are dimensionless parameters that do
not depend on the severity (scale) of ground motions and can be utilized to evaluate relative
strength demands. These reduction factors were discussed in Chapter 3, and the mean and
mean-erR-spectra obtained from the 15 S 1 (15s) records for Jlt
hardening ratios a

= 0.0 and 0.1 are shown in Figs. 3.10 and 3.11.


42

=2

to 8 and strain

Working in the dimensionless R-J..L domain, shown in Fig. 4.8, the modifications
needed for the MDOF base shear capacities, in order to limit the story ductility demands at
the base (assumed to be the maximum) to the prescribed target ductility ratio J..Lr. can be
readily evaluated. For a given number of stories (period) and strain hardening ratio, the
R-J..Lrelationship for any of the three types of MDOF systems and their equivalent SDOF
system can be plotted as shown in Fig. 4.8. Point 1 refers to the strength capacity
associated with an SDOF target ductility ratio f..Lt An MDOF system designed for that
strength capacity will give a higher ductility demand J..L. (shown as Point 2). In order to
limit the MDOF ductility demand to J..Lt, linear interpolation between data points similar to
Point 2 is used to obtain Point 3. The ratio of the reduction factors at Points 1 and 3 is the
reverse of the ratio of the strength capacities of the SDOF and MDOF systems that will give
the same target ductility ratio f..Lt.
Fig. 4.9 illustrates the extent to which the SDOF R-factors need to be modified to
limit the ductility demands in the first stories of the three types of MDOF systems to
specified target ductility ratios. The figure shows the R -J..L relationships for 2 and 40 story
structures with 0 and 10% strain hardening. Data points on which the curves are based are
obtained by using, as ordinates, the mean strength reduction factors for f..Lt = 1, 2, 3, 4, 5,
6, and 8 of the SDOF system with the first mode period of the MDOF systems (these
values defme the strength of both the SDOF and MDOF systems used in the analysis), and,
as abscissa, the mean ductility ratios of the SDOF and MDOF systems. The data points
shown on solid curves for the MDOF systems correspond to Point 2 in Fig. 4.8.
Connecting the data points for each system with straight lines gives approximate R-J..L
relationships for the SDOF as well as MDOF systems. As mentioned before, the ductility
demand parameter used here for the MDOF systems is the story ductility demand in the first
story of the structure, which in most cases is the maximum overall story ductility ratio.
The following observations are made from the R-m relationships in Fig. 4.9:

The R -J..L relationships for the three MDOF models are relatively smooth curves with no
apparent irregularities, justifying the use of linear interpolation between the R -J..L data
points.

The BH model has consistently the closest R-J..L relationships to that of the SDOF
system, especially for strain hardening systems (a= 10% ). Short period BH structures
(2 to 10 stories high) have ductility demands that are even smaller than those predicted
by SDOF systems.

43

I.

The WS model shows the greatest deviation from the SDOF R-JJ. relationships,
particularly for systems with no strain hardening (a= 0%).

The deviation of the MDOF R -JJ. relationships from that of the SDOF system increases
with the number of stories (period). The effect of strain hardening is greatest for the

WS model and smallest for the BH model.


It remains to be answered how the information generated here can be incorporated in
the design process. As was discussed in Chapter 3, the inelastic SDOF strength demand
spectra for specified target ductility values can be derived either from elastic response
spectra and appropriate reduction factors, or may be derived directly from inelastic SDOF
response studies. In order to utilize these SDOF spectra for the design of MDOF systems,
modifications need to be performed. For structures located on S1 soil sites, and which
fulfill the constraints summarized at the beginning of this chapter, the data generated in this
study can be used to evaluate these modifications as illustrated in Fig. 4.8.
Figure 4.10 (parts (a) through (h)) shows the modification factors that need to be
applied to the SDOF strength demand spectra in order to limit the MDOF ductility demand
(in the first story) to the prescribed target values Jl.t = 2, 3, 4 and 8, for 0 and 10% strain
hardening and for the three types of MDOF models used. The period axis comprises the
six different number of stories. The modification factors shown in this figure are the ratios
of the MDOF base shear strength demand Vy(JJ.t) to SDOF inelastic strength demand Fy(JJ.t),
which are the reverse of the ratios of the corresponding reduction factors (see Fig. 4.8), to
limit the ductility demand for both systems to a target value Jl.t
The dashed curves shown in the figure represent the modification factors envisioned
by ATC-3 to account for the multi-mode effect of MDOF systems (that is, raising of the 1!f
tail of the ground motion spectra to I!f2!3 in the design spectra). For soil type S1 and Av =
Aa this ratio, r, is given as :

~ 2.5 Aa
1.2 Av !f
= ---------------213

1.22 Av !f

2.5 Aa

(4.1)

According to the Commentary of the ATC-3-06 document (ATC-3-06, 1978), this ratio is
intended to provide more safety for long period structures and account for higher mode

44

effects. For periods exceeding 0.488 sec. this ratio amounts to 0.984Tl 13. Note that this
modification is independent of the target ductility ratio since ATC-3 assumes a constant
reduction factor, R (or Rw in the 1988 UBC), irrespective of the expected ductility demand
for the given structure.
Fig. 4.11 shows the same information as Fig. 4.1 0, but plotted vs. the target ductility
ratio J.lt instead of period, for 2 and 40 story structures with 0 and 10% strain hardening.
The following observations can be made from Figs. 4.10 and 4.11 :

The required strength modifications for MDOF effects depend strongly on the target
ductility ratio, the fundamental period of the system, the type of structural system, and
the strain hardening ratio a. For BH structures, i.e., structures that involve beam
mechanisms in every story, the modification factor is smallest and is mostly in good
agreement with the ATC-3 modification only if there is significant strain hardening (a=
10%). Even here the caveat exists that, for longer period structures and higher target
ductility ratios, somewhat larger modifications are required than indicated by the ATC-3
approach. For short period BH structures, the base shear strength demand is
consistently lower than the corresponding SDOF strength demand, indicating that
MDOF effects are not important for this range.

For CH structures, i.e., structures in which story mechanisms form within each story
(e.g., plastic hinging in columns, or mechanisms in bracing systems), the MDOF
strength demands are higher than for BH structures. Quantitative information can be
obtained from the graphs. It can be seen that the required increase in strength compared
to the BH structures is about the same regardless of fundamental period. The figures
also clearly demonstrate that WS structures, i.e., structures with a weak first story, are
indeed a great problem. Such structures require strength capacities that may be more
than twice those required in BH structures in order to limit the story drift to the same
target ductility ratio.

A comparison between results for 0 and 10% strain hardening shows that matters get
considerably worse if the structural system has no strain hardening, particularly for CH
structures and longer periods. Systems without strain hardening drift more, and larger
strength is required in order to limit the drift to a prescribed target ductility ratio. This
effect is actually larger than indicated in the graphs, since the SDOF strength demands,
which are used as normalizing factors, are already larger for elastic-plastic SDOF
systems than for strain hardening SDOF systems (see Figs. 3.10 and 3.11). The

45

differences between the results shown for 0 and 10% strain hardening in Figs. 4.10
and 4.11 come only from multi-mode effects.

In general, the modification factor for MDOF vs. SDOF strength demands increases
with target ductility ratios (except for 2-story structures which show an increase then
decrease in modification factors for higher target ductility ratios) and strain hardening.

For the regular structures studied here, elastic MDOF systems attract lower base shears
than that predicted by the equivalent SDOF systems.
These results clearly demonstrate that an estimation of the required strength in the

design process needs to be based on criteria that consider the fundamental period and
distinguish between different "failure" mechanisms and anticipated strain hardening. These
factors may significantly affect the ductility demand the structures will experience in an
earthquake. When drawing quantitative conclusions from the presented data it needs to be
emphasized that the presented results are based on mean values obtained from 15 firm soil
records and, in their final presentation as shown in Figs. 4.10 and 4.11, are derived from
straight line interpolation between discrete ductility values and discrete periods. These
straight line interpolations may lead to considerable approximations for systems for which
the SDOF and MDOF ductility demands differ significantly (e.g., WS structures with high
target ductility ratios).
The foregoing discussion focused on a procedure that can be employed to derive
design strength demands for MDOF systems from inelastic strength demand spectra of

SDOF systems. The presented numerical results apply only within the constraints
identified in this chapter and cannot be generalized without a much more comprehensive
parametric study. The parameters that need to be considered include the frequency content
of the ground motions (which may be greatly affected by local site conditions), the
hysteretic characteristics of the structural models (strain hardening, stiffness degradation,
strength deterioration, etc.), and the dynamic characteristics of the MDOF models (periods,
mode shapes and modal masses of all important modes, as well as stiffness and strength
discontinuities). Furthermore, it must be pointed out that the R-J.l relationships for MDOF
systems developed here are based on ductility demands in the first story. In several cases
the maximum story ductility demand did not occur in the first story, but close to the top of
the structure. This issue is not considered here, but can be addressed through
modifications to the distribution of design story forces over the height for structures

46

without stiffness and strength discontinuities (regular structures), or a case-by-case


dynamic analysis for irregular structures.

4.3 Shear Force and Overturnin2 Moment Demands


The parametric study on MDOF systems provided much statistical information on
shear and overturning moment distribution for elastic as well as inelastic systems. The
maximum story shears and overturning moments for all stories were obtained from the
dynamic time history analyses for the different permutations discussed in Section 4.1.
Depending on the purpose, the maximum dynamic story shears (or overturning moments)
can be normalized in two ways: either by (a) the design base shear (or overturing moment
at the base), to allow a direct comparison between dynamic and design (code) load patterns;
or by (b) the design story shears (or overturning moments), for a story-by-story
comparison.
In all cases, the design story shears (or overturning moments) are the those obtained

from a base shear equal to the strength demand of the corresponding SDOF system and the
1988 UBC seismic load pattern. All results in this section are presented for specific SDOF
target ductility ratios Jlt; the corresponding MDOF story ductility ratios can be obtained
from Figs. 4.4 to 4.6.

4. 3.1 Story Shear Demands


Fig. 4.12 shows an example of the mean a variation of dynamic story shear force
distribution, normalized with respect to the design story shears, for elastic 20-story BH
structures subjected to the 15s ground motions. Figure 4.13 shows the mean distribution
of the elastic dynamic story shears, normalized with respect to the design story shears, for
2 to 40 story structures. The figure shows that the dynamic story shears are somewhat
higher than assumed by code distribution in the upper stories (for tall structures only) and
lower in the lower stories. The values of less than 1.0 at the base indicate that the MDOF
systems attracted lower base shears than their equivalent SDOF systems.
The picture is quite different for inelastic systems. Figure 4.14 shows the same results as
Fig. 4.13 but for inelastic systems with 0% strain hardening and SDOF target ductility ratio
Jlt

= 2.

Figure 4.15 is for 0% strain hardening and SDOF target ductility ratio Jlt

47

= 8 and

Fig. 4.16 is for 10% strain hardening and SDOF target ductility ratio Jlt = 8. The following
can be noted from Figs. 4.14 to 4.16:

For CH structures all stories yield, resulting in uniform ratios of 1.0 over the entire
height of the structures for elastic-perfectly plastic systems (a= 0%), or ratios of 1.0 +
a (p,i- 1) at story i for systems with strain hardening ratio a (see Figs. 4.4 to 4.6). At
any given story, once the dynamic story shear reaches its yield capacity, the story
yields and develops a story mechanism, thus preventing any further transfer of shear
forces across this story unless there is some strain hardening. This is evident by
comparing Fig. 4.15 with 4.16 parts (b) for CH structures for 0 and 10% strain
hardening.

For BH structures, the ratios are always greater than those of the CH structures and
greater than 1.0 even for a= 0%. This is due to the inability of the BH model to
develop a story mechanism, thus attracting more story shear forces. For a= 0%, the
BH model can attract an additionall0-20% story shears due to this effect for Jlr

= 2 and

up to 100% more for Jlt = 8. Note that this would not occur if the BH model were
incrementally loaded with the 1988 UBC load pattern, since a complete structure
mechanism would develop.

The large story shear forces attracted above the first weak story of the WS structures
are illustrated in parts (c) of Figs. 4.14 to 4.16. In contrast to the CH model, making
the stories above the first strong enough to remain elastic attracts much larger story
shear forces, especially in the upper stories. These large shear forces may not pose a
problem by themselves, since for these structures the cause of the weak story is
excessive shear strengths in all upper stories, however, as will be discussed later, they
generate very large overturning moments that have to be resisted in the first story. The
first story will have to resist high axial forces on its columns, besides having to
dissipate most of the seismic energy imparted to the structure. The P-delta effect may
become of major concern for these structures. It is worth noting that the maximum
story shears in the upper stories can reach up to 2 (or 6) times the shears obtained from
the 1988 UBC load pattern, for 0% strain hardening in the first story and Jlt = 2 (<;>r 8),
and even more for 10% strain hardening.

Figures 4.17 and 4.18 show shear forces for 40-story structures with SDOF target
ductilitiy ratios of 2 and 8, normalized with respect to the design base shears. The thin

48

lines refer to the UBC load pattern normalized to a value of 1.0 at the base. The following
observations can be made from these figures :

For inelastic BH and CH structures, the governing dynamic load pattern is closer to a
rectangular (from the almost constant slope shear force diagram) rather than an inverted
triangular shape. This is attributed to the full development of story shear capacities as
illustrated in Fig. 4.14 to 4.16. On the other hand, the WS structures can be
characterized by a 3-step load pattern with maximum intensity at the upper stories and
minimum at the lower stories.

The deviation of dynamic story shears from the design values increases with strain
hardening, target ductility ratio, and towards the bottom of the structures. The
deviation greatly depends on the structural system. It is largest for the WS structures
and smallest for the CH structures.

4. 3. 2 Story Overturning Moment Demands


The issue of overturning moments is considered one of the most critical ones, as
these moments control the flexural design of shear walls and the maximum axial forces that
have to be resisted in columns of braced and moment resisting frames. Since columns
under axial loads are usually brittle, conservative criteria have to be used in estimating
overturning moments, and reductions in overturning moments have to be treated with great
caution.
The shear force distributions discussed previously cannot be used directly to draw
conclusions on overturning moments, because these forces represent maxima in each story
that occur at different times. Therefore, in the dynamic time history analyses, the
maximum overturning moment in each story was computed and stored for evaluation.
Figure 4.19 shows the mean u variation for the dynamic story overturning
moments, normalized with respect to design overturning moments at each story, for elastic
20-story BH structures. The design overturning moments are the overturning moments
obtained from the 1988 UBC load pattern with the base shear set equal to that of the
equivalent SDOF system. It is interesting to note that the variation decreases towards the
bottom of the structure. This is attributed to the cumulative nature of the overturning
moments, which are the summation of the story shears, which are, in turn, the summation
of story forces.

49

Figure 4.20 shows the mean distribution of overturning moments, normalized with
respect to design overturning moments at each story, for 2 to 40-story elastic MDOF
systems. For elastic systems, the moments are close to the design values at the upper
stories (usually not of much importance) and significantly smaller at the lower stories. This
may justify the use of overturning moment reduction factors. However, for elastic
systems, it needs to be emphasized that there are two reasons for the relatively small values
shown at the base of all structures. First, is that the base shears attracted by the elastic
MDOF systems are smaller than the design base shears (see Fig. 4.13), which were based
on the corresponding SDOF systems. Second, the dynamic overturning moments are
smaller than those based on the code seismic load pattern. The contribution of the latter can
be estimated by comparing Figs. 4.13 and 4.20. For example, the mean base shear of
40-story structures is about 0.94 the design base shear, and the mean overturning moment
at the base is about 0.72 the design overturning moment. Thus, the actual reduction in
overturning moments is 0.72/0.94

= 0.77.

For inelastic systems, a much discussed code design issue is whether it is necessary
to design structures for overturning moments that are produced by the story shear
capacities, or whether it is justifiable to use reduced overturning moments assuming that
not all stories yield simultaneously.
Figures 4.21 to 4.23 show the maximum dynamic story overturning moments for an
SDOF target ductility ratio Jlt

for Jlt = 8 and a

= 2 and strain hardening a= 0%, for Jlt = 8 and a= 0%, and

10%, respectively.

The information presented in these figures

correspond to that on shear forces in Figs. 4.14 to 4.18. The following observations can
be made from Figs. 4.21 to 4.23:

The presented results provide much evidence that for structures whose story strengths
are tuned to the design level (all BH and CH structures), simultaneous yielding in most
stories is vezy likely and. therefore. no reductions in overturning moments should be
applied. In the presented graphs, a value of 1.0 implies no overturning moment
reduction, and, as the figures show, this value is reached or exceeded in most cases.
For systems without strain hardening, the CH structures develop the maximum design
overturing moments throughout the entire height of the structure (computed values
close to 1.0), and the BH structures develop even higher demands in the upper stories,
which increases with the target ductility Jlt and strain hardening a. For systems with
strain hardening, the dynamic overturning moments at the base of the structure may be
in the order of 50% higher than design values, for both the BH and CH structures.

50

For inelastic weak story (WS) structures, the dynamic overturning moments become
very large compared to the design overturning moments. For these types of structures,
the ductility demands are very high (Figs. 4.4 to 4.6), the story shear forces are much
larger than the shear strength of the first story (Figs. 4.14 to 4.16), and the overturning
moments exceeds those based on first story strength and code seismic shear distribution
by a large margin (Figs. 4.21 to 4.23). All these issues need to be considered in the
design of weak story structures.
The results presented here give a clear indication that overturning moments in inelastic

structures can be very large. If the story strengths are tuned to the code seismic load
pattern and base shears equal to that predicted by the equivalent SDOF systems for a
prescribed target ductility ratio (as was done in all BH and CH structures), it should be
assumed that all stories will yield simultaneously and the maximum overturning moments
should be based on the shear strengths of all stories above with due consideration given to
strain hardening. No overturning moment reduction factors should be applied. In most
real designs, the story strengths cannot be tuned exactly to the code seismic load pattern
and individual stories may have a shear strength larger than required. In such cases the
overturning moments will further increase. This study provides no information on the
magnitude of this increase, as it depends on the relative strength of each story and cannot
be generalized.
Extreme strength discontinuities, such as those in the WS structures, should be
avoided, whenever possible, as they lead to excessive ductility and overturning moment
demands that may be greatly amplified by the portions of the structure that remain elastic.

51

5.

SUMMARY AND CONCLUSIONS

In the context of long range research objectives directed towards improvements of


seismic design practice, the essential aspects of this study can be summarized as follows:
Consistent protection against failure requires that ductility capacity, which may be
modified for cumulative damage effects, becomes the basic design parameter.
The required strength of structures is a dependent quantity that depends on acceptable
ductility capacities of the important elements of the lateral load resisting system.
Story ductility capacities need to be derived from element ductility capacities in order to
evaluate the required structure (story) strength that "assures" that story ductility
capacities exceed the story ductility demands with an adequate margin of safety.
Story ductility demands, and therefore story strength demands, depend on ground
motion parameters (e.g., severity and faulting mechanism of the eanhquake, source-tosite distance and geologic medium, site soil conditions, etc.) and structural response
parameters (e.g., type of hysteretic system, fundamental period, multi-mode effects,
type of "failure" mechanism, and distribution of strength and stiffness in plan and
elevation). Only regular structures are considered in this study.
For SDOF systems the required strength for specified target ductilities can be evaluated
from statistical studies of ground motion response of representative hysteretic systems,
and can be represented in terms of inelastic strength demand spectra or spectra of
strength reduction factors. Such a statistical study can be attempted only for ground
motions with similar frequency characteristics, such as rock and firm soil motions
recorded not too close and not too far from the fault rupture.
For motions that are greatly affected by local soil conditions, the required strength for
specified target ductilities is strongly site dependent. Site-specific inelastic strength
demand spectra need to be determined; elastic site-specific response spectra can provide
misleading information on inelastic strength demands. Much more research is needed to
assess the effects of site soil conditions on the required strength of structures.
For MDOF systems the required base shear strength must be modified compared to the

SDOF inelastic strength demand in order to account for higher mode effects. A

52

procedure for implementing this modification for regular structures is discussed in this
report. Quantitative conclusions are drawn on the effects of different "failure"
mechanisms on the required structure strength.
The assessment of overturning moments for MDOF systems has shown that in inelastic
systems simultaneous yielding of all stories is a very likely event if the strength
distribution over the height follows the code prescribed load pattern. No overturning
moment reduction factors should be applied, except for structures that respond elastically
in an earthquake. For structures with a weak story (WS structures), the overturning
moments may be much larger than anticipated from the strength of the weak story.

6.

ACKNOWLEDGEMENTS

The work summarized here was in part supported by a grant given by Kajima
Corporation and administered by CUREe, the California Universities for Research in
Earthquake Engineering. Additional support was provided by the John A. Blume
Earthquake Engineering Center at Stanford University, and the Stanford/USGS Institute
for Research in Earthquake Engineering and Seismology. The support from all these
sources is gratefully acknowledged.

53

REFERENCES
ATC 3-06. (1978). "Tentative Provisions for the Development of Seismic Regulations for
Buildings," Applied Technology Council, June 1978.
Boore, D.M., Seekins, L., and Joyner, W.B. (1989). "Peak Accelerations from the 17
October 1989 Lorna Prieta Earthquake," Seismological Research Letters, SSA, Vol. 60,
No.4, October-December 1989.
CDMG (1989). "CSMIP Strong-Motion Records from the Santa Cruz Mountains (Lorna
Prieta), California Earthquake of 17 October 1989," California Department of
Conservation, Division of Mines and Geology, Office of Strong Motion Studies, Report
No. OSMS 89-06.
CDMG (1990). "CSMIP Processed Strong-Motion Records from the Santa Cruz
Mountains (Lorna Prieta) Earthquake of 17 October 1989," California Department of
Conservation, Division of Mines and Geology, Office of Strong Motion Studies, Report
No. OSMS 90-05, August 1990.
Chung, Y.S., Meyer, C., and Shinozuka, M. (1987) "Seismic Damage Assessment of
Reinforced Concrete Members," Report NCEER-87-0022, National Center for Earthquake
Engineering Research, State University of New York at Buffalo, October 1987.

Competing Against Time, (1990). The Governor's Board of Inquiry on the 1989 Lorna
Prieta Earthquake, George W. Housner, Chairman, May 1990.
ldriss, I.M. (1990). "Response of Soft Soil Sites During Earthquakes," Proceedings, a
memorial symposium to honor Professor H.B. Seed, University of California, Berkeley,
May 1990.
Joyner, W.B., and Boore, D.M. (1988). "Measurement, Characterization, and Prediction
of Strong Ground Motion," Proceedings of Earthquake Engineering & Soil Dynamics II,
GT Div/ASCE, Park City, Utah, June 27-30, 1988.
Krawinkler, H., and Nassar, A. (1990). "Strength and Ductility Demands for SDOF and
MDOF Systems Subjected to Whittier Narrows Earthquake Ground Motions," SMIP90,
Proceedings of the Seminar on Seismological and Engineering Implications of Recent
Strong-Motion Data, California Department of Conservation, Sacramento, California, June
8, 1990.
Krawinkler, H., and Nassar, A. ( 1990-1 ). "Damage Potential of Whittier Narrows Ground
Motions," preliminary report submitted to California Department of Conservation, April
1990.
Krawinkler, H., Zohrei, M., Lashkari, B., Cofie, N ., and Hadidi, H. (1983).
"Recommendations for Experimental Studies on the Seismic Behavior of Steel Components
and Materials," John A. Blume Earthquake Engineering Center, Report No. 61,
Department of Civil Engineering, Stanford University.
Mahin, S.A., and Bertero, V.V. (1975). "An Evaluation of Some Methods for Predicting
the Seismic Behavior of Reinforced Concrete Buildings," EERC Report 75/5, Earthquake
Engineering Research Center, University of California, Berkeley, 1975.

54

McCann, M.W., and Shah, H.C. (1979). "Determining Strong-Motion Duration of


Earthquakes," Bull. of Seismological Society of America, Vol. 69, No.4, August 1979.
Nassar, A., and Krawinkler, H. (1991). "Seismic Demands for SDOF and MDOF
Systems," John A. Blume Earthquake Engineering Center Report No. 95, Department of
Civil Engineering, Stanford University, June 1991.
Nims, D.K., Miranda, E., Aiken, I.D., Whittaker, A.S., and Bertero, V.V. (1989).
"Collapse of the Cypress Street Viaduct as a Result of the Lorna Prieta Earthquake," EERC
Report 89/16, Earthquake Engineering Research Center, University of California,
Berkeley, November 1989.
Miranda, E., and Bertero, V. (1991). "Evaluation of the Failure of the Cypress Viaduct in
the Lorna Prieta Earthquake," BSSA Special Issue on the Loma Prieta Earthquake, October
1991.
Osteraas, J.D., and Krawinkler, H. (1990). "Strength and Ductility Considerations in
Seismic Design," John A. Blume Earthquake Engineering Center, Report No. 90,
Department of Civil Engineering, Stanford University.
USGS (1989). ''U.S. Geological Survey Strong-Motion Records from the Northern
California (Lorna Prieta) Earthquake of October 17, 1989," USGS Open File Report No.
89-568, October 1989.
USGS (1990). "Lorna Prieta, California, Earthquake of October 17 1989 - Processed
Strong-Motion Records- Volume 1," USGS Open File Report No. 90-xxx, March 1990.

55

Table 3.1 Specific Hysteresis Models Generated from the General Model Shown in Fig. 3.1
SKELETON PARAMETEilS

MODEL

Fy

Fs

HYSTERESIS PARAMETERS

Ke

Kt

Kh

Ke

Kt

Kh

Fy

Fs

Ke

Basic Bilinear

Ke

Kh

Fy

Bilinear with Strength Det.

Ke

Kh

Fy

Bilinear with Unloading


Stiffness Degradation

Ke

Kh

l'y

Clough Model
basic Stiffness Degradation

Ke

Kh

ry

Ke

Clough Model with Ace.


Stiffness Degradation

Ke

Kh

l'y

Clough Model with


Strength Deterioration

Ke

Kh

Fy

Clough Model with


Unloading Stiff. Degradation

Ke

Kh

Basic Pinching Model

Ke

Pinching and Accelerated


Stiffness Degradation

Ke

Pinching and Strength


Degradation

Ke

'

Basic Trilinear

Ku

Kr

... j)

B,

Kp

... d

Ke

Ke

Ke

Ke

Ke

ps rd 0
'

Rule

omax

Ke

Rule

I\ 0 to
.

Ke

Rule

Omax

l'y

Pu Ku,o

Rule

Omax

Kh

Fy

Ke

Rule

Rule

Rule

Om ax

Kh

Fy

Ke

Rule

Rule

Rule

1\ ot.()

Kh

Fy

Ke

Rule

Hule

Rule

0 max

Pu K11,0 puKu,o

ps rd 0

ps rd,o

Table 3.2 Properties of the 15 S 1 Records from Western U.S. Eanhquakes Used in Statistical Study

175.9

17:7

392.4

30.0

156.8

20.5

392.4

30.0

.5

392.4
30.0
30.0
30.0

VI
~

6.

27.0

5.6

192.7

6.3

5.

12.0

6.5

39.1

PGV obtained by numerical integration of digitized acceleration records

6.4

29.3

392.4

59.

15.8

392.4

39.7

296.2

30.0

44.0

392.4

70.1

167.9

30.0

32.3

364.5

.0

228.1

24.9

178.7

14.7

392.4
392.4

Table 3.3. Lorna Prieta Rock Site Records Used in this Study

GROUND MOTIONS ON ROCK SITES


Record

Name

Comp.

Number
57007
47379
58065
1601
58127
47189
47377
58219
58130
58338
58151
58163
58133
58131
58132
58043

Corrialitos ( Canyon Road)


Gilroy #1 (Gavilan Coli.)
Santa Cruz (UCSC/ Elec. Lab.)
Stanford Univ. (Slac Test Lab.)
Woodside (Fire Station)
Sago South (Hoi. Cienegaa Rd.)
Monterey (City Hall)
Hayward (CSUH Stadium)
SF. (Dimond Heights)
Piedmont (Jr. High Grounds)
SF. (Rincon Hill)
Yerba Island (Yerba Buena)
SF. (Telegraph Hill)
SF. (Pacific Heights)
SF. (Cliff House)
Ptbonita (Point Bonita)

0
90
0
360
90
261
0
90
90
45
90
90
90
270
90
297-

Distance (km)
PGA (cm/sec. 11 2) PGV (em/sec.)
Epicenter to Rupture
7
1
618
55.2
29
15
434
33.8
16
16
433
21.2
51
33
282
28.4
55
36
14.7
80
54
39
71
10.3
44
49
69
3.3
71
53
7.4
83
73
92
111
14.3
74
93
81
9.2
76
95
89
11.6
77
95
89
11.6
78
97
91
9.6
78
97
14.3
60
99
80
106
21.0
104
85
71
13.6

Table 3.4. Lorna Prieta Alluvium Site Records Used in this Study

GROUND MOTIONS ON ALLUVIUM SITES


Record
Number
58065
47125
47380
47381
57066
57425
1656
57191
57382
47179
57064
1686
58393
58498
58505

Name

Comp.

Saratoga (Aloha Ave.)


Capitola (Fire Station)
Gilroy #2 (Hwy 101/Bolsa Rd.)
Gilroy #3 (Sewage Plant)
Agnew (State Hospital)
Gilroy #7 {Mantelli Ranch)
Hollister (Differential Array)
Halls Valley (Grant Park)
Gilroy #4 (San Ysidro School)
Salinas (John and Work St.)
Fremont (Fire Station)
Fremont (Emerson Coun)
Hayward (Muir School)
Hayward (Bart Station)
Richmond (City Hall Pkg. Lot)

0
0
0
0
0
90
255
0
0
250
0
90
0
310
280

Distance (km)
Epicenter to Rupture
27
9
14
9
16
30
18
31
40
25
40
28
30
45
31
37
32
32
46
34
55
39
56
40
71
53
73
55
108
89

58

PGA (cm/sec."2) PGV (em/sec.)


494
463
344
532
163
314
281
128
408
110
118
191
166
155
123

41.3
36.1
33.3
34.5
30.9
16.3
36.6
12.5
39.1
15.8
10.2
10.8
13.6
11.8
17.1

Table 3.5. Other Lorna Prieta Records (Mostly Soft Soil Sites)

TABLE OF OTHER RECORDS


Record
Number
47006
57476
57383
1695
47524
1002 .
58375
58223
58539
58472
58472
58222
58117
1662
58471
1678
68003

Name

Comp.

Gilroy (Gavilan Coli. Phys.)


Gilroy (2-Story Historic Bldg.)
Gilroy #6 (San Ysidro)
Sunnyvale (Colton Avenue)
Hollister (South & Pine)
Redwood City (A peel Array 2)
Foster City (Redwood Shores)
SF. (International Airport)
South SF. (Sierra Pt..)
Oakland (2-Story Office Bldg.)
Oakland (Outer Harbor Wharf)
SF. (Presidio)
Treasure Island (Fire Station)
Emeryville ( Christie Ave. )
Berkeley (Lawrence Lab)
Golden Gate (Abtmant Bldg.)
Olema (Ranger Station)

67
90
90
360
0
43
90
90
205
290
35
90
90
260
90
270
0

Distance (km)
PGA (em/sec. "2) PGV (em/sec.)
Epicenter to Rupture
29
15
349
28.9
28
16
280
43.5
24
35
167
13.9
43
215
33.4
48
362
33
62.8
272
63
53.1
44
63
278
45.4
79
60
326
29.3
84
65
103
8.2
92
73
238
37.9
76
281
95
40.8
79
98
195
33.5
79
98
156
33.4
97
255
41.1
114
99
80
22.0
100
239
35.5
138
119
158
18.5

59

Table 3.6. Motions Recorded at or Near Stanford

Name

Source

Ep. Dist.
(km)

PGA
(cm/sec2)

PGV
(em/sec)

282

28

Stanford, SLAC (SLAC)

USGS #1601

51

Stanford, Parkg Garage (SPG)

USGS

51

,-_.).)

33

Palo Alto VA Hospital (V APA)

USGS #1227

47

378

Menlo Park VA Hospital (VAMP)

USGS #1230

54

288

.:+0
24

Table 3.7. Typical Soft Soil Records


Ep. Dist.

Source

Name

(km)

PGV
Site Cond.
PGA
(cm/sec2) (em/sec)

CDM.G #47524

48

362

63

"alluvium

Foster City, Redwood Shores CDMG#58375

63

278

45

"alluvium"

SFO, Int. Airport

CDMG#58223

79

326

29

"alluvium"

Oakland, 2-story Office Bldg. CDMG#58224

92

238

38

"alluvium''

Emeryville, 6363 Christie Av. USGS #1662

97

255

41

fill on mud

USGS #1601

51

282

28

sed. rock

Hollister, South Str. & Pine

Stanford, SLAC

Table 3.8. Japanese Records Used in this Study

TABLE OF JAPAN'S RECORDS


Record Name
Tokyo 101 NS
Osaka 205 EW
Nagoya 306 NS
Sendai 111-030 EW
Sendai 111-038 EW
Sendai 501 EW
Hachinohe NS
Hachinohe EW

Record

Magnitude

. Date
2.14.1956
3.27.1963
3.27.1963
6.12.1978
6.12.1978
4.30.1962
5.16.1968
5.16.1968

5.9
6.9
6.9
7.4
7.4
6.5
7.9
7.9

Unsealed
PGA cm/sec."2

74.00
19.00
10.00
258.23
240.90
45.00
224.95
182.93

60

Scaled to PGV:SO em/sec.


PGVcm/sec.

6.08
2.23
1.29
36.15
36.32
4.73
33.27
40.43

PGA cm/sec.":Z

608.55
426.00
387.59
357.16
331.63
475.68
338.07
226.23

PG V em/sec.

50.00
50.00
:5~J.oo

50.00
50.00
50.00
50.00
50.00

Table 4.1 Modal Periods and %Mass for MDOF Structures Used in this Study
2-STORY
Mode
#

1
2
3
4
5

(sec)

Mass
%

0.217
0.089

90
10

5-STORY

10-STORY

20-STORY

30-STORY

40-STORY

(sec)

Mass
%

(sec)

Mass
%

(sec)

Mass
%

(sec)

Mass
%

(sec)

Mass
%

0.431
0.176
0.111
0.082
0.064

82
11
4
2
1

0.725
0.288
0.181
0.133
0.106

80
11
4
2
1

1.220
0.475
0.294
0.214
0.169

78
11
4
2
1

1.653
0.636
0.391
0.282
0.221

78
11
4
2
1

2.051
0.781
0.479
0.345
0.269

78
11
4
2
1

61

SEAOC LOCAL DUCTILITY DEMAND


72'x120'x12', a= 0.0, S = 1.2
10

MOMENT FRAME

PERIMETER FRAME
8

BRACED FRAME

----- ------

0 0----------------------------------------------~
1
4
2
3
ELASTIC PERIOD (seconds)

Fig. 2.1. Estimates of Local Ductility Demands for Code Designed Steel Frame Structures
Designed According to the 1988 UBC (SEAOC Approach)
(from Osteraas and Krawinkler, 1990)

62

DAMAGE THRESHOLD

COLLAPSE THRESHOLD

Identify acceptable
deflection and strength
reduction factors

Identify member
ductility capacities

Local
ductility
capacities

Global
ductility
capacity

t
Global
ductility
capacity

Global
strength
demand
I

Derive local
strength and stiffness
demands

'

Global
strength
demand

Local
strength
demand

/
Compare local strength demands
Design tor controlling condition

Revise design

!
Static, nonlinear analysis

I
Revise assumptions

with targets
and assumptions

Final design

Fig. 2.2 Flow Chart of Ductility and Drift Based Dual Design Approach
(from Osteraas and Krawinkler, 1990)

63

RESISTANCE

Fig. 2.3 Basic Seismic Demand Parameters

RESISTANCE

DISPL

Fig. 2.4 Plastic Deformation Range Used for Damage Evaluation

64

GENERAL HYSTERETIC MODEL

Ku

I
I

r
L

Fs

Ke = Elastic Stiffness
Kt =Transition Stiffness
Kh =Hardening Stiffness
F y = Yield Strength
F s = Structure Strength
Ku =Unloading Stiffness
Kr =Reloading Stiffness
Kp =Pinching Stiffness
F d =Yield Strength after Strength Deterioration
F p = Pinching Target Resistance

ot

= Target Displacement

Fig. 3.1 Parameters of GeneralHysteresis Model

65

SPEC.

..
E

Bt-<;

i.'?

12 00

__/

--...J

~.)7,)9

-I

L
0

..

e.co<~

DETERIORATION
TIIRESIIOLD

4.8"

-"1

r'

s
.00

1--1

E-<

V1
H
V1

- ee

(a)

-e .ee_

-1.59

c .....

I.

5"

EN!:' DEF't.EC:TION IN

J.eCI

NUHBER OF CYCLES

(a) Response of Test Specimen

(b) Mode of Deterioration

Fig. 3.2 Test Specimen with Large Deterioration Threshold

-c

.
E

12-1

12 . .

..

...

IC

4 ..

II

I.
0

D.T

DETERIOHATTON

-~r

!I

.ee

-4 . . .

(b)

- ..
_,

....

-a ...

NlJNI.IER OF CYCJ.lS

(a) Response of Test Specimen

(b) Mode of Deterioration

Fig. 3.3 Test Specimen with Small Deterioration Threshold

66

DET.

Bilinear Mode/with Strength Deterioration


a'== lOO C::l.O

(b) Deterioration in Yield StrengrJ, With N lllnber of Cycles

Fig. 3.4 StrengrJ, Deterioration in Bilinear HYsteresis MOdel


(a::::: 100 and c::::: 1.0)

67

Bilinear Model with Strength Deterioration


a= 100

-2

-4

c:2

-l

i
-0.5

(a) Hysteresis Model

Deterioration of Yield Strength


a

= 100

c:2.0

1.2

1.0

0.8
-o
u.

0.6
0.4
0.2
0.0
0

10

Number of Cycles

(b) Deterioration in Yield Strength with Number of Cycles


Fig. 3.5 Strength Deterioration in Bilinear Hysteresis Model
(a= 100 and c = 2.0)

68

12

Bilinear Model with Degradation of Unloading Stiffness


0.=100

C:1

(a) Hysteresis Model for a= 100 and c = 1.0

Bilinear Model with Degradation of Unloading Stiffness


0. = 100

C:2.0

(b) Hysteresis Model for a= 100 and c = 2.0


Fig. 3.6 Degradation on Unloading Stiffness in Bilinear Hysteresis Model

69

Clough Model with Accelerated Stiffness Degradation


0.=100

C:1.0

(a) With Accelerated Stiffness Degradation

Clough Model with Strength Deterioration


Ct

=100

C:1.0

I
I

i
1 ~

-4

(b) With Strength Deterioration


Fig. 3.7 Deteriorating Peak Oriented Models Subjected to Constant Amplitude Cycling
(a= 100 and c = 1.0)

70

Clough Model with Accelerated Stiffness Degradation


a=100

c:1.0

(a) With Accelerated Stiffness Degradation

Clough Model with Strength Deterioration


a =100

c:1.0

1.5 ..;..

(b) With Strength Deterioration


Fig. 3.8 Deteriorating Peak Oriented Models Subjected to Cycling with Different Amplitudes
(a= 100 and c = 1.0)
71

STRENGTH CAPACITY vs. DUCTILITY DEMAND- {bi-10)


Alhambra (270), Bilinear, T = 0.9 sec, a= 10%
4 ,-------------------------------------------------.

3 ,__ _ _ _ _ _ _~~----~~---------~

CIS

E
Q)

>.

~2-+---------------------=:::......,.,..-------l
:l

1 ~.--------------~-----------------------------~

0.00

0.04

0.08

0.12

Strength Capacity, Fy(Jl) I W

(a) Nonrnonotonic Relationship Between Fy(J.L) and Jl

DISPLACEMENT TIME HISTORIES FOR DIFFERENT STRENGTH CAPACITIES (bi-3.1 0)


Alhambra (270), Bilinear, T = 0.9 sec, Jl == 3, a= 10%

Time {sec)

(b) Displacement Time Histories for the Three SDOF Systems with J.L- 3 in Fig. 3.9(a)
Fig. 3.9 Problems in the Determination of Inelastic Strength Demands for Given J.L
(from Krawinkler and Nassar, 1991)

72

STRENGTH REDUCTION FACTOR, Ry(J.l.)- (15s.bi-OO)


15s Records, Bilinear, a= 0%, Mean

12
10
8

......

'-'>-

a:

6
4

~ =2, 3. 4. s. 6. 8 (thin --+thick lines) ~

0.0

0.5

1.0

1.5

2.5

2.0

3.0

3.5

4.0

T (sec)
(a) Mean Values

STRENGTH REDUCTION FACTOR, Ry(J.l.)- (15s.bi-OO)


15s Records, Bilinear, a= 0%, Mean-a
8 ~----------~----------~----~----~----~--~

~ =2, 3, 4, 5, 6, 8 (thin --+thick lines) -

~~----~----~----_j----~-===~==========~====~
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

. T {sec)
(b) Means minus Standard Deviation
Fig. 3.10 Strength Reduction Factors for Bilinear Systems with no Strain Hardening

73

STRENGTH REDUCTION. FACTOR, Ry(Jl)- (15s.bi-10)


15s Records, Bilinear, ex = l 0%, Mean
12
10
8

.::;.6

a:

2
J.l = 2, 3. 4, 5, 6. 8 (thin

~thick lines) ~

0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

T (sec)

(a) Mean Values

STRENGTH REDUCTION FACTOR, Ry(J.L)- (15s.bi-1 0)


8

15s Records, Bilinear, ex= 10%, Mean-<1


I

.-----~----~----~-----,----~~---,-----------,

!
I

J.l =2. 3, 4, 5, 6, 8 (thin

~thick lines) ~

0~,----~--~----+----+~==~===r====r===~
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

T (sec)
(b) Means minus Standard Deviation
Fig. 3.11 Strength Reduction Factors for Bilinear Systems with 10% Strain Hard~ning

74

NORMALIZED HYSTERETIC ENERGY


strain hardening

=0%

- mean

80 ~----~--~----~--~~--~,----~----~--~
)1='2.3,J,5,6,8 (thin -thick :ines)

30 ~----~----~----------~----~!----~----~----~

>>-

:.c

~ JO ~~~~~--~----~----~----~----~--~~--~
UJ

-....
0.0

1.0

0.5

1.5

2.5

2.0

3.0

4.0

3.5

T (sec)

(a) Mean Values

NORMALIZED HYSTERETIC ENERGY


strain hardening
120
100

A ,._
40
20

1'4 \
V\

V'-

I
I

- --

0
0.0

.........

j--1

1-

~-

,.___ I

1.0

I
l

0.5

- mean+<r

)P2.3,4,5,6,8 (thin- thick lines)

i-.

:-I
~
_\j ~ ~""- ~
./""'k
'oj

=0%

1.5

2.0

2.5

'

'

.
I

'

3.0

'
I

3.5

4.0

T (sec)

(b) Means plus Standard Deviation


Fig. 3.12 Norm. Hysteretic Energy Demands for Bilinear Systems with no Strain Hardg

75

NORMALIZED HYSTERETIC ENERGY


strain hardening

=10%

- mean

80 ~--------------------------------------~~-------~----~--------~

0.0

0.5

, .0 .

, .5

2.0

2.5

3.0

3.5

4.0

T (sec)

(a) Mean Values

NORMALIZED HYSTERETIC ENERGY


strain hardening = 10% - mean+cr
120 .--------~----~----------~-----~~----------~-----------------------~
p2.3.4.S.6.8 (thin- <hick lines)

100 ,_~~~~-----+--------+-----~----~----~----------~----~

o~~~~~==~==t===C===~=====d
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

T (Set?)
(b) Means plus Standard Deviation
Fig. 3.13 Norm. Hysteretic Energy Demands for Bilinear Systems with 10% Strain Hardg

76

NUMBER OF INELASTIC EXCURSIONS, N


strain hardening

=0%

- mean

40 ,-----~----~----~----~----.-----~--------~
~o<=2.3.4..S.6.S

(thin- r.uck lines)

!
i

30 ~-F;-------------------~----~----~----~--~

0 ~.----~----~----------~----~----------~--~
0.0

0.5

1.0

.5

2.5

2.0

3.0

3.5

4.0

T (sec)

(a) Mean Values

NUMBER OF INELASTIC EXCURSIONS, N


;

strain hardening

= 0%

- mean+cr

60 ,-----~----~--------~~----~----~--~----~
~2.3.4.5.6.8

(thin- thick lines)

50 ~----7-----r-----~--~----~'----~----~----~

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

T (sec)

(b) Means plus Standard Deviation


Fig. 3.14 Number of Inelastic Excursions for Bilinear Systems with no Strain Hardg

77

NUMBER OF INELASTIC EXCURSIONS, N


strain hardening

= 10%

- mean

40 ~--------------------~----~,----~--------~
~2.3.-1.5.6.3

(thin- :hick lines)

0 ~~----~----~----------~----~--------~----~
4.0
3.5
3:J
2.5
2.0
1.5
1.0
0.5
0.0

T (sec)

(a) Mean Values

NUMBER OF INELASTIC EXCURSIONS, N


strain hardening

=10%

- mean+cr

60 .-----~--------------~----~----~--------~
~2.3,4,5,6,8

(thin -!hick lines)

50 ,_H+~~----~----~---+----~----~----~--~

0 ~----~----~--------~----~----~----~--~
1.5
0.0
1.0
4.0
0.5
2.0
2.5
3.0
3.5

T (sec)

(b) Means plus Standard Deviation

Fig. 3.15 Number of Inelastic Excursions for Bilinear Systems with 10% Strain Hardg

78

MEAN OF THE PLASTIC DEFORMATION RANGES


strain hardening = 0% - mean

5 ~----------~~----------~~-----~~----------~----------~----~--~----------~
J.l"'2.3,4,5,6,3 (thin- thick lines)

4+---~'--~~--~----r---7---~---+--~
~3 +--------~--~--~~~--~-------~----~~~====~

:.c

0~~----------~----------------~-----~----------~-------~-------~--~

0.0

, .0

0.5

1.5

2.0

2.5

3.0

3.5

4.0

T (sec}

(a) Systems with noStrain Hardening

MEAN OF THE PLASTIC DEFORMATION RANGES


strain hardening

=10%

- mean

5 ~----------~~----------~~----------~----------~----------~----------~-----~-------~
!J"'2.3,4.S,6.8 (tl'lin- illicit lines)

--r--1

~I

~r--r----~;;:;;;;;;;;:;=:

~3+---~1--~~~~~~==~~+-~

:.c

--

~~2~~~~~~--~~~--~~~f=~:-r---~====~

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0.

T (sec}

(b) Systems with 10% Strain Hardening


Fig. 3.16 Mean Values of the Normalized Plastic Deformation Ranges

79

HYSTERETIC I TOTAL DISSIPATED ENERGY

=0%

strain hardening
1.0

UJ
Q

,..,.

';

y~

~ -

0.5

.....
UJ
.,...

/""',

OA

I
'

'

'

0.2

i'
\

I!

!I

''

ii

i
i

f!=2.3A.5.6.3 (th:n- th:ck lines)

!
!
0.0

I
I

I
i

0.8

iI

- mean

iI
;

3.0

3.5

'

0.0

0.5

1.Q

1.5

2.0

2.5

4.J

T (sec)

(a) Systems with no Strain Hardening

HYSTERETIC I TOTAL DISSIPATED ENERGY


strain hardening
1.0

I
i

0.8

0.5

;-

:
;'

UJ

::r: 0.4

../

0.2

i
I

'

,-

I
I

~i

.....

~2.3.4.5.6.8

(thin -thick lines)

II

'

0.0

0.5

I
1.0

1.5

I
I

I
I

2.0

2.5

3.0

3.5

0.0

!~

I
I

J'V

- mean

'
~

UJ
Q

= 10%

4.0

T (sec)

(b) Systems with 10% Strain Hardening


Fig. 3.17 Contribution of Hysteretic Energy to Total Dissipated Energy

80

EFFECT OF STRAIN HARDENING ON A-FACTORS - (15s.bi-OO/bi-1 0)


15s Records, Bilinear, (a= 0%) I (a= 10%), Mean

1.4

-....
0

.!.
.0

i
I

1.2

!
'

1.0

.......

.....

;"'

0.6

0.4

;
I
i

>o

0.2
0.0

J..L

i'

::::::::...::::.

'

!
I

= 2, 3, 4, 5, 6, 8 (thin ~ thick lines) ~

;
I
I

i'

iI

II

2.0

2.5

3.0

3.5

I
'

~
~
I

II

I!

i!

!
I

!I

""'

'-'

a:

~.....

.!.
.0

~-!~

";? 0.8

-g

0.5

0.0

1.0

1.5

4.0

T (sec)

Fig. 3.18 Effect of Strain Hardening on Strength Reduction Factor

RATIOS OF HYSTERETIC ENERGIES - (15s.bi-00/15s.bi-10)


15s records, bilinear, (a;::O%)/(a=lO%)
1.2

-,..

l..o-.

1.0

-"'. K:J
,,r-

~ 0.8

0.6

s.
w

0.4

I
I

'(

w
:::c:
-

I
I

I J.I"'2.3.4.S.6.8 (thin -

:z:::

0.2

0.5

1.0

1.5

2.0

2.5

3.0

3.5

0.0
0.0

ltlic:k lines)

4.0

T (sec}

Fig. 3.19 Effect of Strain Hardening on Hysteretic Energy Dissipation

81

ELASTIC & INELA~IC STRENGTH DEMAND SPECTRA (15s-OO)


Scaled to PGA = 0.4 (g) - mean

1.2 ,---------__.;.;:~---;::iiiiiiiiiiiiiiiiiiiiiiiiiiiiiiiiiiil
--ATC-5 1

--,.1
--,.=2
--,.3
--,..6

1.0

--j.ia4

--J.l"S

0.8

J.1.=8

Cl

""-;.. 0.6

u..

0.4
0.2
0.0

J.---:-.~.:===~~~~~~~~~;~~~~~;~
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

T (sec)

(a) Mean Values

ELASTIC & INELASTIC STRENGTH DEMAND SPECTRA (15s-OO}


Scaled to PGA

1.6

=0.4 (g) - mean + a


ATC-5 1

1.4

--J.ll

1.2

--1'2
--J.1.3
--J.L-4
--J.LS
--J.L-6

1.0

,. .. s

C)

""-;.. 0.8

u..

0.8
0.4
0.2
0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

35

4.0

T (sec)

(b) Means plus Standard Deviation


Fig. 3.20 Strength Demand Spectra of Records Scaled to PGA = 0.4g (no Strain Hardg)

82

ELASTIC & INELASTIC STRENGTH DEMAND SPECTRA. (15s10)


Scaled to PGA

=0.4 (g) - mean

1.2 r---------........:.::~-----;iiiiiiiiiiiiiiiii;iiiiiiiiiiiii;l
- - ATC-S 1
--~,.1

1.0

--~,.2

--~ .. 3
--~ .. 4

--~

0.8

.. s

--~=6
~=8

C)

-;.. 0.6

u.

0.4

0.2

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

T (sec)

(a) Mean Values

ELASTIC & INELASTIC STRENGTH DEMAND SPECTRA (15s-10)


1.6

Scaled to PGA =0.4 (g)- mean+ a


T________
. ;.; .;.___

~~~ffi

ATC-S 1
--~o~l

1.4

--1''"2
--~o~3

1.2

--~o~a4.

--~o~aS

--~o~-6

1.0

~o~-8

-;.. 0.8
LL.

0.8
0.4

0.2~~~~~
0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

T (sec)

(b) Means plus Standard Deviation


Fig. 3.21 Strength Demand Spectra of Records Scaled to PGA

=0.4g

(10% Strain Hard:: 1

83

..

ELASTIC & INELASTIC STRENGTH DEMAND SPECTRA - (15s-OO)


1.2

Scaled to PGV =12 in/sec - mean

.----------------;iiiiiiiiiiiiiiiiiiiiiiiiiiiiiiiJ
--ATC.S,
--~&1

1.0

--~&2
--~&3
--~&4

--~&S

0.8

--~&6

).l8

CD

-;. 0.6

.....

0.4
0.2

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

-4.0

T (sec)

(a) Mean Values

ELASTIC & INELASTIC STRENGTH DEMAND SPECTRA - (15s-OO)


Scaled to PGV = 12 in/sec - mean + a
1.6 r------------~;iiiiiiiiiiiiiiiiiiiiiiiiiiiiiiil
--ATC.S1

1.4

--~&1

--~&2
--~&3

1.2

--~&4

--~&S

1.0

--~&6

J.LI

CD

"";. 0.8

.....

0.8

0.4

0.0

05

1.0

1.5

2.0

2.5

3.0

3.5

-4.0

T (sec)

(b) Means plus Standard Deviation


Fig. 3.22 Strength Demand Spectra of Records Scaled to PGV = 30 em/sec. (no Strain H.)

84

EFFECT OF HYSTERESIS MODEL ON A-FACTORS- (15s.dg-10/bi-10)


15s Records, Degrading I Bilinear, ex= 10%, Mean
1.4

-.....
0

I
)\

.0

:::::: 1.0

'

"'h.

--

0.8

I
i

C)

:::::: 0.4

:::::t.
>-

a:

0.2
0.0

l.---::
I

i
i

0.5

1.0

1.5

2.0

2.5

0.0

I
I

1.

I J.1 =2, 3, 4, S. 6, 8 (thin -+ thick lines) ~

I
!

'0

II

~ 0.6

-vi

I
I

k,

.!..

rl"

1.2

II

!I

3.0

3.5

4.0

T (sec)

Fig. 3.23 Effect of Stiffness Degradation on Strength Reduction Factor

EFFECT OF HYSTERESIS MODEL ON NHE - (15s.dg-OO/bi-OO)


15s Records, Degrading I Bilinear, ex= 0%, Mean
5

04
0

J.1 =2, 3, 4,
;I

I
I

iI

.!..
.0

'

w3

-:1:

9C) 2

-z
'0

:1:

I
I

I:

I
I

5, 6, 8 (thin -+thick lines) ~

1\

Ii

~~-!

0.0

0.5

1.0

1.5

2.0

2.5

3.0

iI
I

I
I

3.5

4.0

T (sec)

Fig. 3.24 Effect of Stiffness Degradation on Hysteretic Energy Dissipation

85

ATTENUATION OF PGA
700

Cll

<

"

a..

'b0

u
0

500

e
Cll

400

I
\

300

''

200

' ' .... ....

100

... ......

<

"a..

..---------~

00

Joyner & Boo"' Prediction

600

Regression Curve (h-e! km)

Alluvium Sites

700

Joyner & Boore Prediction

600

....
e

ATTENUATION OF PGA

Rock Sites

\
\

400

I
I

''

300
200

''

' ' ....,,

100

Regression Curve (h-a i<m)

500

...........---

--- --- -----------

04-----~------~-----.------~-----4
o
~
~
oo
ao
100

Horizontal Distance to Rupture Surface (km)

Horizontal Distance to Rupture Surface (km)

(a) Rock Site Records

(b) Alluvium Site Records

Fig. 3.25 Attenuation of Peak Ground Acceleration

ATTENUATION OF PGA
Alluvium & Rock Sites

700

Alluvium Sites
Roc:!< S~es

600

'b0
Ill

500

400

<

300

"
a..

200

----- -----

100

so

40

eo

100

Horizontal Distance to Rupture Surface (km)

Fig. 3.26 Attenuation Relationships Derived from Rock and Alluvium Site Records

86

A TIENUATION OF ELASTIC STRENGTH DEMAND SPECTRA


Rock Sites

1.4

r0 = 10 km
r0 = 20 km

1.2

r0 =40 km
r0 = 80 km

1.0

SMF88

CMF68

0.8

.9

.......

a)

:>:

u..

.......

0.6

.............

...... ....... .......

0.4

--

0.2

ZC for S 1 (1988 USC)

--- --- --- ----

--------------------

0.

0.

0.5

1.0

1.5

2.0

T {sec)

(a) Rock Site Records

A TIENUATION OF ELASTIC STRENGTH DEMAND SPECTRA


Alluvium Sites

1.4

r0

.9

1.2
1.0

r0 = 40 km
r0 =80 km

""
"

0.8

a)

:>:

u..

= 10 km
= 20 km

r0

SMF88
.......

CMF68
.......

.......

......

ZC for S2 (1988 UBC)

......

0.6

.................. -... .........

0.4
0.2

-- -- -- --- ---

--------

0.

0.

0.5

1.0

1.5

T {sec)

(b) Alluvium Site Records


Fig. 3.27. Attenuation of Elastic Strength Demand Spectra

87

2.0

VARIATION OF DYNAMIC AMPLIFICATION FACTORS


4.0

Rock Sites Elastic Systems


-r-------------=--_;_
_______,
r0 -10km

ro a 20 km
ro a 40 km
3.0

u.
ct

r0 80km

2.0

1.0

0.

0.5

1.0

1.5

2.0

T (sec)

(a) Rock Site Records

VARIATION OF DYNAMIC AMPLIFICATION FACTORS


4.0 _,....._ _ _ _ _ _
___
__
__
.:..__ _ _ _ _ ___,
Alluvium
Sites
Elastic
Systems
r0 10km
r0 20 km
r0 40 km
r0 80 km

3.0

u.
ct
Q

2.0

1.0

0.

0.5

1.0

1.5

2.0

T (sec)

(b) Alluvium Site Records


Fig. 3.28 Variation of Shapes of Elastic Strength Demand Spectra with Distance

88

ATTENUATION OF INELASTIC STRENGTH DEMAND SPECTRA


Rock Sites -

0.6

1.1.

=4

I
I

I
I

I
I

0.5

r0 10km
r0 20 km
r0 40 km
r0 80 km

I
I

I
I

--

0.4

I
I

......,

.:;

''

0.3

''

::...

""

' ' ,,,,

SMF88
CMF68
',, ,,

0.2

...........

0.1

,,

-- -----------------------------

--------------------------0.
0.

0.5

1.0

1.5

2.0

T (sec)

(a) Rock Site Records

ATTENUATION OF INELASTIC STRENGTH DEMAND SPECTRA


Alluvium Sites - 1.1.

0.6

=4

I
\

0.5

r0 10km
r0 20km
r0 40km
r0 SOkm
SMF88
CMF68

I
I
I
I

--

I
I

0.4

......,

.:;

''

0.3

::...

""

''

''

--,

-,,

,,

--

0.2

------ ......... _
-----------------------

0.1

--------------0.
0.

0.5

1.0

1.5

2.0

T (sec)

(b) Alluvium Site Records


Fig. 3.29. Attenuation of Inelastic Strength Demand Spectra for J..L =4

89

VARIATION OF NORMALIZED STRENGTH DEMAND SPECTRA


Rock Sites - (Normalized w.r.t. PGA) - ll = 4

1.0

r0 -10km
r0 20 km
r0 .. 40 km
r0 80 km

"tJ

c
ca
E
Q)

0.8

-...

.c
en
c
Q)

-"2

0.6

UJ

0.4

.!::!
16

...0

0.2

0.
0.

1.0

0.5

1.5

2.0

T (sec)

(a) Rock Site Records

VARIATION OF NORMALIZED STRENGTH DEMAND SPECTRA


Alluvium Sites - (Normalized w.r.t. PGA) - ll = 4
1.0

r0 10km
r0 20km
r0 40km
r0 80km

"C

c
ca
E
Q)

0.8

.c
a
cQ)

-i...

en

0.6

0.4

iii

...E
0

0.2

0.
0.

0.5

1.0

1.5

2.0

T (sec)

(b) Alluvium Site Records


Fig. 3.30 Variation of Shapes of Inelastic Strength Demand Spectra (J.L =4) with Distance

90

REGRESSED ELASTIC AND INELASTIC STRENGTH DEMANDS


Rock Sites e

1.4

=10 km
J.l=l
J.1=2
J.1=3
J.l=4

1.2
1.0

-s

SMF88
CMF68

0.8

G)

LL.

0.6
0.4
0.2
0.
0.

1.5

1.0

0.5

2.0

T (sec)

(a) Rock Site Records

REGRESSED ELASTIC AND INELASTIC STRENGTH DEMANDS


Alluvium Sites e

1.4

J.l=l
J.1=2
- - J.l=3
J.l=4

1.2
1.0

--

0.8

LL.

0.6

C)

=10 km

SMF88
CMF68

CD

0.4

0.2

-----------0.

0.

0.5

1.0

1.5

2.0

T (sec)

(b) Alluvium Site Records


Fig. 3.31 Regressed Elastic and Inelastic Strength Demand Spectra at ro = 10 km

91

ELASTIC RESPONS.E SPECTRA

1.4

-r--------------------------,
-

SLAC

_sPG
1.2

--VAMP
- - VAPA

1.0

0.8

~
II.

0.8

0.4

0.2

o.

0.5

1.0

1.5

2.0

T (sec)

Fig. 3.32 Elastic Response Spectra of Motions Recorded at or Near Stanford University

92

STRENGTH DEMAND SPECTRA

STRENGTH DEMAND SPECTRA

1.5 ,...._ _ _
___
_ _Str.
__
__...,;(e
__
_ _ _,
Hollister,
South
& Pine
48_
km)

Stanford, SLAC (e =51 km)


1.5 - , - - - - - - - - - - - - - _ _ ;_ _ _ _...,

1'=1
1'=2
1'=3
1'=4

ZC (UBC 1988)

1.0

).'=I
ZC (USC 1988)

5 2 54

-----"\--~----,

1.0

5MF88

', ',

' ', ' '..........

CMF68

,.

',

,.

.........

u.

u.

',,

........

f'=4

... .... .............

__

CMF68

_ ,____ ----- --

.... ....
..........................

....

0.5

0.0

5MF88

.....

_____ ------

--- ---

----------------

+--------:-----:----~~-----~

0.0

0.5

1.0

2.0

1.5

T (sec)

T (sec)

STRENGTH DEMAND SPECTRA

STRENGTH DEMAND SPECTRA

1.5 ...,..._ _Foster


_ _ _City,
......;._Redwood
_ _ _ _Shores
___
_ _ _,
(e_=_
63_
km)

ZC (UBC 1968)

1.5 ...,..._ _ _
_ _International
_ _ _ _ _Airport
_...;__;....;.,
...:.,._ _..,
SFO,
(e_ _
79 _
km)

1'=1
1'=2
1'=3
1'=4

52 54

1.0

5MF88
CMF68

0.5

1.5

2.0

T (sec)

T (sec)

STRENGTH DEMAND SPECTRA

STRENGTH DEMAND SPECTRA

1.5 ...,..._ _Oakland,


____
_..;...Office
___
...;....;...
_km)
_...;_ _.,
2-story
Bldg.
(e _92

1.5

Emeryville,
6363
Christie
Av._(e_=_
97_km)
___
__
___
_ _ _,

....--~-.....;

I'= I

1'=2
1'=3
1'=4
1.0

5MF88
CMF68

,.

u.

0.5

0.5

1.0

1.5

2.0

T (sec)

T (sec)

Fig. 3.33 Strength Demand Spectra for Six Lorna Prieta Ground Motions

93

VARIATION OF SHAPE OF ELASTIC SPECTRA


Elastic Systems
5.0,.-------------------EI'Tl8fyville
Foster City
Hollister
Oakland
SFO Airport
Stanford

~4.0

Q.

" '0

; 3.0

-...

.::: 2.0
CD

cC),

U) 1.0

0.0 _,__ _ __...,-._..._...,.....,....or--.,...._ _ _,.......,,.......,........

_~

0.0

0.5

1.0

1.5

2.0

T (sec)
Fig. 3.34 Normalized Elastic Strength Demand Spectra

STRENGTH REDUCTION FACTORS


Foster City, Redwood Shores
10.0 . . . , . . . . . - - - - - - - - - - - - - - - - - - - .
)1=4
)1=3
)1=2

8.0

6.0

4.0

2.0

0.0

-+------r--------,r-------,..----.
..
1.5
2.0
0.5
1.0

0.0

T (sec)
Fig. 3.35 Strength Reduction Factors for Foster City I Redwood Shores Record

94

VARIATION OF SHAPE OF INELASTIC SPECTRA

).1.=4
1.0 . , - - - - : - - - - - - . . . . . ; . . - - - - - - - - - -....

<
(!)

Emeryville
Foster City
Hollister
Oakland
SFO Airport
Stanford

0.8

.-c, 0.6
c.
ca

ID
Q

-..
-

.c 0.4
Cl
c
ID

...

tn 0.2

0.0

'

-t------r----.....-...,-.....-.....-.....-.....-...,-.....-.....-.....-....---.

0.0

0.5

1.0

1.5

2.0

T (sec)
Fig. 3.36 Normalized Inelastic Strength Demand Spectra (for f.1 =4)

STRENGTHDEMANDSPECTRA
stanford, SLAC (e =51 km)

1.5,-----------.....;..---~----~
Corrected
Hand Digitized

1.0

T (sec)
Fig. 3.37 Spectra Obtained from Hand-Digitized and Corrected Stanford/SLAC Record

95

STRENGTH CAPACITY E2 FOR GENERIC STRUCTURES


Concrete & Steel Structures
0.6
SBF88

0.5

SMF88

--

SPF88

0.4

>-

CMF68

C)

0.3

CMF88

\
\

u.

0.2

,.
' ..... , ........ ...... -- ... ,
......... ......... ... .... ..._
.... ____

. -----

0.1

-----==~~~---~

0.0
0.0

1.0

0.5

1.5

2.0

T (sec)
Fig. 3.38 Strength Capacity of Five Types of Code Designed Building Structures

96

SITE SPECIFIC STRENGTH DEMAND SPECTRA


1.5

Tokyo
(NS) _ _ _ _ _ _ ____,
_101
_.....;..___;_

-.-:--~-----___;;

j..L=l
j..L=2
j..L=3
j..L=4

, .0

j..L=5
j..L=6

-Ci

j..L=8

>.

0.5

0.

JL~~~~_:~~~~~~~~~~~~~~~;;~~
0.

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

T (sec)

(a) Tokyo 101 NS Record

SITE SPECIFIC STRENGTH DEMAND SPECTRA


Hachinohe (NS)

, .5 --..------------.....;....___;__ _ _ _ _ _ ____,
j..L=l
j..L=2
j..L=3
j..L=4

1-L=S

1.0

j..L=6

c;

j..L=8

>.

0.5

0.

0.5

, .0

1.5

2.0

2.5

3.0

3.5

4.0

T (sec)

(b) Hachinohe NS Record


Fig. 3.39 Elastic and Inelastic Strength Demand Spectra (Bilinear Systems, a= 0.1)

97

SITE SPECIFIC STRENGTH DEMAND SPECTRA


Elastic Systems J.L=l

2.5 -r---------......!.---=-_.:;_~-------.-,
Hachinohe EW
Hachinohe NS
2.0
Nagoya 306 NS
Osaka 205 EW
Sendai 030 EW
1.5
Sendai 038 EW
Sendai 501 EW
Tokyo 101 NS
u.>1.0

-s

0.5

0.
0.

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

T (sec)

(a) Elastic Systems

SITE SPECIFIC STRENGTH DEMAND SPECTRA


Inelastic Systems J.1.=2

1.0-r--------------------~----~----------------~

Hachinohe EW
Hachinohe NS
Nagoya 306 NS
Osaka 205 EW

Sendai 030 EW

c;

.._..

u?

Sendai 038 EW
Sendai 501 EW
Tokyo 101 NS

0.5

0.

0.5

1.0

1.5

2.0

2.5

3.0

3.5

T (sec)

(b) Inelastic Systems with f.L = 2


Fig. 3.40 Strength Demand Spectra for 8 Japanese Records

98

4.0

SITE SPECIFIC STRENGTH DEMAND SPECTRA


Inelastic Systems J.1=3

1.0-r--------------------~----~-----------------

C)

.._.

Hachinohe EW
Hachinohe NS
Nagoya 306 NS
Osaka 205 EW
Sendai 030 EW
Sendai 038 EW
Sendai 501 EW
Tokyo 101 NS

0.5

u.?

0. -LI~~~~~~~::~~====~::~~~~~~~;;~~
0.

0.5

1.0

2.0'

1.5

2.5

3.0

3.5

4.0

T (sec)

(c) Inelastic Systems with Jl. = 3

SITE SPECIFIC STRENGTH DEMAND SPECTRA


Inelastic Systems J.1=4

1.0~--------------------~----~----------------~

Hachinohe EW
Hachinohe NS
Nagoya 306 NS
Osaka 205 EW
Sendai 030 EW
Sendai 038 EW
Sendai 501 EW
Tokyo 101 NS

0.

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

T (sec)

(d) Inelastic Systems with J1. = 4


Fig. 3.40 (cont'd) Strength Demand Spectra for 8 Japanese Records

99

NORMALIZED HYSTERETIC ENERGY


Inelastic Systems J.l. = 2

30.0-r--------------------~~----~--------------~

Hachinohe EW
Hachinohe NS
Nagoya 306 NS
Osaka 205 EW
Sendai 030 EW
Sendai 038 EW
Sendai 501 EW
Tokyo 101 NS

20.0

c.o>>-

u.
UJ
::z:::

10.0

0.
0.

0.5

2.0

1.0

2.5

3.0

3.5

4.0

T (sec)

(c) Inelastic Systems with J.J. =2

NORMALIZED HYSTERETIC ENERGY


Inelastic Systems J.l.

70.0

=4
Hachinohe EW
Hachinohe NS
Nagoya 306 NS
Osaka 205 EW
Sendai 030 EW
Sendai 038 EW
Sendai 501 EW
Tokyo 101 NS

60.0
50.0
. >>- 40.0

c.o

u.
UJ

J:

30.0
20.0
10.0
0.

o.

0.5

1.0

1.5

2.0

2.5

3.0

3.5

T (sec)

(d) Inelastic Systems with J.J. = 4


Fig. 3.41 Normalized Hysteretic Energy Dissipation Demands

100

4.0

HYSTERETIC ENERGY
Inelastic Systems J.L = 2

25000.0

J\

20000.0

( ,)

Q)

15000.0

tn

(,)

UJ
J:

Hachinohe EW
Hachinohe NS
Nagoya 306 NS
Osaka 205 EW
Sendai 030 EW
Sendai 038 EW
Sendai 501 EW
Tokyo 101 NS

10000.0

I
/'\

5000.0

I\
I \
I \
I \
I
\
I
\

0.
0.

1.0

1.5

2.0

2.5

3.0

3.5

4.0

T (sec)

(c) Inelastic Systems with J.l. =2

HYSTERETIC ENERGY
Inelastic Systems J.L = 4

20000.0-r------------------~----~---------------,

Hachinohe EW
Hachinohe NS
Nagoya 306 NS
Osaka 205 EW
Sendai 030 EW
Sendai 038 EW
Sendai 501 EW
Tokyo 101 NS

15000.0

~
-

( ,)
Q)

10000.0

(,)

UJ
J:

5000.0

0.

0.5

1.0

1.5

2.0

2.5

3.0

3.5

T (sec)

(d) Inelastic Systems with J.l. = 4


Fig. 3.42 Hysteretic Energy Dissipation Demands per Unit Mass

101

4.0

.___ \..- kp

--y-

--- ---

(a) Basic Bilinear Model

Qy

-.. .a..............

..... ....
I

........

(b) Peak Oriented Degrading Bilinear Model


Fig. 3.43 Hysteresis Models Used for Study in Section 3.6

102

100

~. ~~~~~~~ll---'-.~
<

lOlJ

(a) Time History

Cm/S e

c:)

,--------,--------~------~

20

l 5

10

5. 0

1. 0

2. 0

3. 0

PERIOO<SECl

(b) Fourier Amplitude Spectrum


(c. 1)
1 500

.--------,.-----,-------,

1000

500

0
0

1. 0

2. 0

PERIOD<SECJ
(c) Elastic Response Specrtum

Fig. 3.44 El Centro Record


103

3. 0

100

u
u
<

-HI.J

(a) Time History

(Cm/Sec)

20

.----------.----------~--------~

1~

10

5. 0

PER[OD<SECJ

(b) Fourier Amplitude Spectrum

c CoIl

I 5 00

r-----------.,-----------,-----------,

1000

500

0
0

2. 0

1. 0

PER[OD<SECJ

(c) Elastic Response Specrtum


Fig. 3.45 Hachinohe Record
104

3. 0

'"

,,
I

(a) Time History


(Y)t!:ae htstory
tCm/Sec)
20 ~------~--------~--------

l5

l0

PERlOD<SECl

(b) Fourier Amplitude Spectrum


(Cal)

1 500

!000

~
500

lr~

r--.._

l. 0

2. 0

PERlOD<SECl

(c) Elastic Response Specnum


Fig. 3.46 Taft Record
105

--

3. 0

..:
X

w
0

z
w
'-'
<
<

:::E
0

.;

"'-:,,:::;::::~

ba:uc b.i linear 1model

''

1.0

o.'

J.O

NA7URAL PER!OD<SECl

Fig. 3.47 Damage Indices (El Centro)

peak Ot"ientedl

degrading type

..:
X

"'z
0

"''-'<
:::E

<

.;

basic bilinear aodel

0. 0

I 0

' 0

NATURAL PER!OD<SECl

Fig. 3.48 Damage Indices (Hachinohe)

106

J. 0

I
I

-z
w

'-'
<

::0:

<
c

o.o

Fig. 3.49 Damage Indices (Taft)

107

l 0

(a) Beam Hinge (BH) Model

(c) Weak Story {WS) Model

(b) Column Hinge (CH) Model

Fig. 4.1 Types of Structures Used in MDOF Study

Vy (J.1) =Modified MDOF Base Shear Capacity


for a target ductility ratio llt

Strength Capacity

').(J.'t) = SDOF Strength Capacity


for a target ductility ratio J..lt

-----;r------------~~

____ _,

I
I
I
I
I

SDOFI

MDOF

I
I
I

. "-1

Target Value J.Lt

1
J.L'

Ductility Ratio, ~

Fig. 4.2 Modification in MDOF Strength Capacity to Achieve Equal


Ductility Demands in SDOF and MDOF Systems
108

DYNAMIC STORY DUCTILITY DEMANDS, Jli- (15s.bh20-8.00)


15s Records, BH Model, 20-Story, llt = 8, a= 0%, Meancr

1.0 --;-----;:-;r---:-,---:------r----...;...,----~
'

:1
j
0.8 -+---:...,.._....___)rf----;-1!- - - - ; - - - - - + - - - - - - + - - - - - 1

~
C1)

0.6

J:

'-l

CD

.2:

c;

-+--___;,.--t-+----t---~-----+------+-------1

0.4

G)
a:

.,

-~

L11

+-------!.--,---f-loo-lr--'-o--~-----+-----+-------1

--:-~'-:-:

!._:

0.2

'--:
'---,
'
:
-+-----+-~-:-:__--'_-;r-'\ - . -__
-- - - - + - - - - - - - - ! - - - - - - - 1
''

''

c\ :}_ '-t"j

0.0 -+-----+---_.__----:,..____.__ _--+------+--------1

10

15

20

25

Story Ductility Ratio, J.li = odyn,l' Oy,i


(a) BHModel

DYNAMIC STORY DUCTILITY DEMANDS, Jli -

15s Records, CH Model, 20-Story, Jlt = 8, a= 0%, Meancr

1.0

''

~ \

t--:

'

'

':

0.8

r-

'

C)

J:

:- ~~--J
'

.c
G)

(15s.ch20-8~00)

0.6

CD

>
:;::

as 0.4

G)

a:

(.

'

'

.~~LL
~

'

0.2

_,

'

.
~'-1
'

I
:

:
:

I
:

0.0
0

10

..:

15

20

25

Story Ductility Ratio, J.li = 8dyn,11 Oy,1


(b) CH Model
Fig. 4.3 Variation in Story Ductility Demands for 20-Story Strucrures for J.lt(SOOF) = 8, a= 0% (15s Records).

109

0.8

C)

0.6

Q)

>

2,5,10,20.30,40 stories
(thick -+thin lines)

~ 0.4
Q)

a:

0.0 +._ _ _ _ _ ___........

0.2

_ _ _ _ _ _ ___.l

=odyn,1/ oy,l

Story Ductility Ratto, Jli

(a) BH Model
DYNAMIC STORY DUCTILITY DEMANDS, Jli- (15s.ch-2.00)

1.0

0.8

C)

Q) 0.6

:t:

Q)

>

2,5,10,20,30,40 stories
(thick-+ thin lines)

~ 0.4
Q)

c::

0.2

0.0

Story Ductllity Ratio, Jli

=odyn,l/ oy,l

(b) CH Model
DYNAMIC STORY DUCTILITY DEMANDS, Jli - (15s.ws-2.00)

1.0

~~

=
~

0.8

C)

0.6

c;

o.4

Q)

2,5,10,20,30,40 stories
(thick -+ thin lines)

a:

0.2

_I

_l_

0.0
0

10

12

Story Ductility Ratio, Jli = odyn,l I Oy,l

(c) WS Model
Fig. 4.4 Dynamic Story Ductility Demands for J.lt(SDOF) = 2, a= 0% (Mean for 15s.Recordsl

110

1.0

DYNAMIC STORY DUCTILITY DEMANDS, J.li (15s.bh-8.00}

-r-~~r-rr----------__;...:.......;._

_ _ _.....:...,

0.8

.C)
::
~

0.6

Cl)

>

2,5,1 0,20,30,40 stories


(ttlick-+ ttlin lines)

~ 0.4

a;

a:
0.2
0.0 +.---_,....-U........L...-J.----1.-,.....;;.;;l.--~----,-----1
5
10
15
20
25
0
30

Story Ductility Ratio, J.li

=odyn,i I Oy,i

(a) BH Model
DYNAMIC STORY DUCTILITY DEMANDS, J.li (15s.ch8.00)

1.0

0.8

. ::

C)

0.6

Cl)

>

2,5, 10,20,30,40 stories


(ttlick -+ ttlin lines)

~ 0.4

a;

a:
0.2

0.0
5

10

15

Story Ductility Ratio, J.li

25

20

30

=odyn,l/ oy,l

(b) CHModel
DYNAMIC STORY DUCTILITY DEMANDS, J.li (15s.ws-8.00)
1.0

-!

h:
i)

0.8

.C)
::

0.6

CD

>

~ 0.4

a;

2,5, 10,20,30,40 stories


(ttlick-+ ttlin lines)

a:

0.2

0.0
0

10

20

30

40

Story Ductility Ratio, J.li

50

60

70

80

=Sdyn,1t ay,1

(c) WS Model

Fig.4.5 Dynamic Story Ductility Demands for J.lt(SDOF) = 8, a= 0% (Mean for 15s Records)

111

1.0

DYNAMIC STORY DUCTILITY DEMANDS, J..1.i (15s.bh-8.1 0)


........,...----------.......:;..!-_;__ _ __:_,

~~-r--r-

0.8

.:
0

0.6

Cl)

>

2,5,10.20,30,40 stories
(thick-+ thin lines)

~ 0.4

Qi

a:

0.2
0.0 +.----....L....I.--.L...-1........=_,..---.....-------.....J
5
10
15
20
25
30
0

Story Ductility Ratio,

J..1.i

= odyn,l I 5y,l

(a) BH Model
DYNAMIC STORY DUCTILITY DEMANDS,

J..1.i

(15s.ch-8.10)

1.0

0.8

.:

C)

G) 0.6

:I:

Cl)

.::

2,5,1 0,20,30,40 stcries


(thick-+ thin lines)

'(; 0.4

Qi

a:

0.2

0.0
5

10

15

25

20

Story Ductility Ratio, J..l.i

30

=odyn,l I Oy,J

(b) CHModel
DYNAMIC STORY DUCTILITY DEMANDS, J..1.i (15s.ws-8.10)
1.0

!/ !;-.

0.8

.:

0.6

Cl)

>

0.4

Qi

a:

0.2

2,5,10,20,30,40 stories
(thick -+thin lines)

II

0.0

10

20

30

40

Story Ductility Ratio,

50
J..1.i

60

70

80

=8dyn,l I 5y,J

(c) WS Model
Fig. 4.6 Dynamic Story Ductility Demands for ~(SDOF) = 8, a= 10% (Mean for 15s Records)

112

FIRST STORY DUCTILITY DEMAND, ~ 1 (15s-2.00)

..-12

r-/v

:1

- - - - SOOF
- e - MOOF- BH Model
-o-- MOOF- CH Model
--e- MDOF- WS Model

-g-10

ca

E
~8
>-

.'1:::

:56
u
::s

i!-4

02

/
v

FIRST STORY DUCTILITY DEMAND, ~t 1 (1552.10)

.-12

r-

::1..

-c~10

- - - - SDOF
--.-- MDOF- BH Model
-o-- MDOF- CH Model
--e- MDOF- WS Model

~8

>-

:1:::

su

::s

..Q

_._____ ...

i!-4

...n---

~'"V"

U) 2

u::

U:::o
0.5

0.0

1.5

1.0

2.5

2.0

---

--

0
0.0

0.5

1.0

T (sec}

(a)

).1 1(SDOF)

=2, a= 0%

::1..

-c
70
c

- - - - SDOF
--.-- MOOF- BH Model
-o-- MOOF - CH Model
- t 1 - MDOF- WS Model

ca

E 60

~50
~

~ 40

.s 30
i!'
0

u;

i?

u:

20

10 ~---

'"'

0
0.0

_/

.--v

(b)

....-

1.5

1.0

T (sec)

2.5

2.0

./

-d' 70
:ij
E 60

f
lO
1..
.

2.0

~t 1 (SDOF)

= 2, a= 10%

FIRST STORY DUCTILITY DEMAND, ~ 1 (1558.10)


:I 80 r,::;~~~~~.---.,-.....:...:....!...-T--~

(1558.00)

--1------- -----~

_.,;;K. _ _ _

--0.5

_1---V

1.5

T (sec)

FIRST STORY DUCTILITY DEMAND, ~ 1

... 80

Lo

- --

f!

L-a---

--

2.5

- - - - SDOF
--.-- MDOF - BH Model
-o-- MDOF- CH Model

SO

- - - MDOF - WS Model

40

----- -----1-------- -----1----

~k~ ----

~~ ~-'~
----=-[~ --=== =---=I.J""\1---~

u:

10

---

--1--- . _;_;_

--

'"'
o~------~----~-------4-------+------~
0.0

0.5

1.0

1.5

2.0

2.5

T (sec)

(c) ).1 1(SDOF) = 8, a= 0%


,
(d) J.1 1(SDOF) = H, a= 10%
Fig . .4.7 Comparison Between First Story Ductility Demands for MDOF Stntctures and SDOF Target Ductility Ratios (Mean for 15s Records)

R-J.J. RELATIONSHIPS FOR


SDOF AND MDOF SYSTEMS
R (J..L)
y

SDOF

Ry(JJ.)(MDOF)

Ry(J..Lt)(SDOF) _ Vy(J..Lt)
Ry(J..Lt)(MDOF) - Fy(J..Lt)

Fy.e.
Vy(IJ.)

Ry(J..Lt}(MDOF) = Fy.e
Vy(JJ.t)
1.0

1.0-

Fig. 4.8 R-JJ. Relationships for SDOF and MDOF Systems (Explanation)

114

STRENGTH REDUCTION FACTORS, R (11) (15s.0200)

1------~---

- - - - SDOF
1 -t----1----+---~-t_._ MDOF- BH Modal
-o- MDOF- CH Model
-11- MDOF- WS Model

- - - - SDOF
+-_.~:....--.-+----+---!---1_._ MDOF- BH Modal
-o- MDOF- CH Model
-11- MDOF- WS Model

o-~--~-----4-----+--~=+====~====~

ot-----~--~-----4--~=+====~==~~
2

......

12

10

12

10

Ductility Ratio, 11

Ductility Ratio, 11

(a) 2 Stories (T =0.217 sec), a= 0%

(b) 2 Stories (T = 0.217 sec), a= 10%

......
Ul

STRENGTH REDUCTION FACTORS, R (11) (15s.4D-10)

STRENGTH REDUCTION FACTORS, R (11) (15s.40-00)


12

12

10 _ _ , __ .
8

:i

~6-~+-~-~--~--~---~~~~l

>,6

a:

a:

10

20

30

40

50

60

70

-+
I
I

- - - - SDOF
__._ MDOF- BH Modal
-o- MDOF- CH Model
-11- MDOF- WS Model

Ductility Ratio, 11

I
I
-II
I

- - - - SDOF
__._ MDOF - BH Modal
-o-- MDOF - CH Model
-a- MDOF- WS Model

80

10

20

30

40

50

60

Ductility Ratio, Jl

(c) 40Stories(T=2.051 sec),a=O%


(d) 40Storics(T=2.051 sec),a= 10%
Fig. 4.9 R-~t Relationships for SDOF and MDOF Systems (Mean for 15s Records)

70

80

BASE SHEAR MODIFICATION FOR MDOF EFFECTS ' VVI(MDOF) IFY'(SDOF) (1552 00)
3.0

- - - ATC-S 1
- - . - MDOF- BH Model
-o- MDOF CH Model
-11- MDOF WS Model

iL 2.5
0
0

,..
u..
U)

2.0

- 1.5

~~
"' ~----

iL

0
0

:IE

1.0

> o.5

BASE SHEAR MODIFICATION FOR MDOF EFFECTS V (MDOF) IF. Y'(SDOF) (155210)

'

3.0

iL

- - - ATC-S 1
- - . - MDOF- BH Model
-o- MDOF- CH Model
-11- MDOF- WS Model

2.5

0
0

--

r---

Cl)

2.0

1.5

;t-

-k>

~ ~------

iL

0
0

:E

~
.,

1.0

lo(

......

> o.5

0.0

_ ..... I"'\...:::::::::: ~

-C

--

0.5

1.0

1.5

2.0

0.5

0.0

2.5

1.0

T (sec)

(a) J.l,

=2, a= 0%
'

3.0

2.5

Cl)

2.0 -11- MDOF- WS Model

:::::. 1.5

"0
0

:IE

""S.

>

- - - ATC-S 1

- e - MDOF- BH Model
-o- MDOF- CH Model

1.0

~
::::: .,

vr- .......... --

2.5

BASE SHEAR MODIFICATION FOR MDOF EFFECTS V (MDOF) IFyl(SDOF) (155310)

-!o

~-- ------~ -------.........

'

3.0

lL

U)

2.5
2.0

::::.. 1.5

"0
0

:E

""S.

>

0.5
0.0
0.0

2.0

(b) J.l, =2, a= lO%

1.5

T (sec)

BASE SHEAR MODIFICATION FOR MDOF EFFECTS V (MDOF) IFY'(SDOF) (15s3 00)

jl_ _____

0.0

0.0

li:'

:--a-

1.0

0.5

- - - - ATC-S 1
MDOF- BH Model
-o- MDOF- CH Model
- - - MDOF- WS Model

-e-

~
..'

.()_

~ ~--- r-------

. -

""' -7"""

/'

~----

0.0

0.5

1.0

1.5

T (sec)
(c) 11, = 3, a= 0%

2.0

2.5

00

0.5

1.5

1.0

T (sec)

(d) ~1 1

= 3, a= 10%

Fig. 4.10 Base Shear Modification for MDOF Effects vs. Period ( 15s Records)

2.0

2.5

BASE SHEAR MODIFICATION FOR MDOF EFFECTSI V


v(MDOF) IF, V'(SDOF) (1554 00)
6

- - - - ATC-S1

(LS

(LS

MDOF- CH Model
__.,_ MDOF WS Model

(/) 4

-3

:e

>.
lL

1--

LL'

;>-1

___...

-~ __.-------- -----,.._-- fa

_,..,_---

~-

- - - - ATC-S 1
- . - MOOF - BH Model
-o- MDOF- CH Model
- - - - MDOF- WS Model

--+- MDOF BH Model

-o-

0
cn4

BASE SHEAR MODIFICATION FOR MDOF EFFECTS I V


"Y'(MDOF) IF, Y'(SDOF) (155410)

._ 3

......

lL

0
0

:E

>-

>1

....

0.0

0.5

1.5

1.0

.....

2.0

2.5

0.5

0.0

T (sec)

(e) llt =4, a= 0%

(0 J1 1 = 4, a= 10%

- - - - ATC-S 1

G:'5
0

--+- MDOF BH Model

-o-

MDOF CH Model
__.,_ MDOF WS Model

(f)4

-3

LL'

:e

~1

./

---

- - - - ATC-S 1
MDOF- BH Modol
MDOF CH Modol
- - - - MDOF WS Model

-+-o-

(LS

>.

lL

/
-- ----- - ------

_.--

_.-v-

lL

0
0

---

:E

2 ----

>-

>1

--

::::.3

-0

2.5

2.0

BASE SHEAR MODIFICATION FOR MDOF EFFECTSI V'VI(MDOF) IFyl(SDOF) (155-810)

(/) 4

---v-

....

1.5

1.0

T (sec)

BASE SHEAR MODIFICATION FOR MDOF EFFECTSI V'y1(MDOF) IF'y1(SDOF) (1558 00)

-n--;:

---

......

------ ------ ----

.....

t-

~
'S;;!"' -

-a:

.....

........---- - - - - - - ..

...---~

------

.--

- - - - - ~----

----- -------.- z.::::::::::


-----

0.0

0.5

1.0

1.5

2.0

2.5

0.0

0.5

1.0

1.5

T (sec)
T (sec)
(h) J1 1 = 8, a= 10%
(g) Jl 1 =8, a=O%
hg. 4.10 (Cont'd) Base Shear Modification for MDOF Effects vs. Paiod (15s Records)

2.0

2.5

BASE SHEAR MODIFICATION FOR MDOF EFFECTS I V'y1(MDOF) IFy(SDOF) (15s 0200)
1.5

...

~.----

tL

0
0

U)

~
-..

tL
c

U)

~
.._

'

:e

0.5

~
-a
2

(b) 2 Stories, a= 10%

.------

'

- - - - ATC-S 1
__.__ MDOF- BH Model
-o- MDOF- CH Model
- t 1 - MDOF - WS Model

0
c
tn4

~
-a
i:L

1"\..

c
:e

-~ ...-------- -------- --------

>.

>1 - - -

BASE SHEAR MODIFICATION FOR MDOF EFFECTS, V, Y'(MDOF) IF (SDOF) (155 40.10)
(LS

~-

;:-1

(a) 2 Stories, a= 0%

/
/~

(L

Target Ductility Ratio, Jlt

- - - - ATC-S 1
- - - - MDOF BH Model
-o- MDOF CH Model
- t t - MDOF WS Model

0
c
tn4

Target Ductility Ratio, 1 1

(L6

- - - - ATC-S 1
__.._ MDOF - BH Model
-o- MDOF- CH Model
- - - - MDOF- WS Model

0.0
. 0

BASE SHEAR MODIFICATION FOR MDOF EFFECTS, V'y1(MDOF)/ F (SDOF) (155 40.QO)

:e

F==

c 0.5
:e
>>

- - - - ATC-S 1
__.._ MDOF - BH Model
-o- MDOF- CH Model
----- MDOF- WS Model

0.0

1.0

...

;:.

.--

.........

tL

tL

1.5

>-

1.0

BASE SHEAR MODIFICATION FOR MDOF EFFECTS Vy(MDOF)/ Fy(SDOF) (155 0210)

~
:a

.-

---4

~-

I----

- -- - -

Target Ductility Ratio, 111

Target Ductility Rallo, 11 1


(c) 40 Stories, a= 0%
.
(d) 40 Stories, a== 10%
Fig. 4.11 Base Shear Modification for MDOFEffccts vs. Target Ductility Ratio ~t 1 ( 15s Records)

MAX. DYNAMIC STORY SHEARS, Vdyn,i IVdes,i - (15s.20-e)


15s Records, 20-Story, Elastic, Meancr

1.0

--r--------------:---.-..-...-.. -...-..-. ~~~~::~:::,-.-...-..-...-.. -...-..-.. --....,

0.8

-t----------;---------.,.-'---rY--+--r-:~...-..-:---.:..-.. .,.------1

0.6 - t - - - - - - - - - - - . , . - - - - - - - : - - - - , - -__''__ ,...,._L::-----'----------1

C1)

I
I

0.4

a:

-+-----------+----~~--....:...'-----------1
I
\ :.. ... II

0.2

- t - - - - - - : - - - - - - - : - - - - - : .~1---:---i-~.J- - - - - - - f

0.0

:1
-+~--------------------~~~----------1

0.5

0.0

1.5

1.0

Vdyn,i I Vdes,i
Fig. 4.12 Variation in Maximum Dynamic Story Shears for Elastic 20-Story Structures (15s Records)

MAX. DYNAMIC STORY SHEARS, Vdyn,i I Vdes,i- (15s-e)


15s Records, Elastic, Mean
1.0

0.8

.r:.
C)
C1)

::z::

0.6

f})

lr:f ~

C1)

>
.;

0.4
..5!
C1)

2,5, 10,20,30,40 stories


(thick~ thin lines)

a:

0.2

0.0

II~
1

0.0

0.5

1.0

1.5

Vdyn,i IVdes,i
Fig. 4.13 Maximum Dynamic Story Shears for Elastic Structures (Mean for 15s Records)

119

0.8

.c
0

~ 0.6
Q)

>

2,5,10,20,30.40 stories
(thick --.thin lines)

~ 0.4
Q)

a:
0.2

+-------,------..,.-L.;..-----------..J

0.0

1.0

0.5

0.0

1.5

2.0

V d~n,l IV des,l
(a) BH Model
1.0

MAX. DYNAMIC STORY SHEARS, V dun


1 I V des 1 (15s.ch2.00)

0.8

.c
0

~ 0.6
Q)

>

0.4

Q)

a:

2,5, 10,20,30,40 stories


(thick--. thin lines)
.~

0.2

0.0 +-------,------..&.....----~------!
1.5
1.0
0.5
2.0
0.0

Vdyn,l I Vdes,l

(b) CH Model
MAX. DYNAMIC STORY SHEARS, V d n 1 / V des 1 (15s.ws-2.00)
1.0 ~-------:::F:::t:==f~~~~-----,

c;-

0.8

.c
0

~ 0.6

a;

2,5, 10.20,30,40 stories


(thick --. thin lines)

0.4

a:

0.2

0.0

0.5

1.5

1.0

2.0

3.0

vdyn,l/ vdea,l
(c) WS Model

Fig. 4.14 Max. Dyn. Story Sheazs (Nonnalized by Vdes) for J.Lt(SOOF) = 2, a= 0% (Mean for 15s Records l

120

.C)
=.
~

0.6

>
~ 0.4

41

G)

a:

II

2.5,1 0,20,30,40 stories

\(

(thick

~thin

lines)

0.2

l._____,....____ll.t~=~~--.......--~

0.0

0.0

0.5

1.0

1.5

2.5

2.0

3.0

Vdyn,i I Vdes,l

(a) BH Model
MAX. DYNAMIC STORY SHEARS, Vdvn 1/ Vdes 1- (15s.ch-8.00)

1.0~----------------------------r-------------------~~~~~~-------~

-!

0.8

.=.
C)

0.6

41

>

~ 0.4
Q)

a:

2,5,1 0,20,30,40 storiesj


(thick

thin lines)

0.2
0.0~----------~------------~------------~-------~------------~-------~

0.0

0.5

1~

15

2~

2.5

3.0

Vdyn,l/ Vdes,l
(b) CH Model
MAX. DYNAMIC STORY SHEARS, V n 11 Vdes 1- (15s.ws-8.00)
1.0 1--,--.-'::::::l;e~;;;;:=~~--=-=~~----,

0.8

.=.
C)

~ 0.6
41

>

a;

2,5, 10,20,30,40 stories

0.4

(thick

&!

~thin

lines)

0.2
0.0 +.--E=::;=:::=::::::::!..._..------.-----,----~
0
2
4
6
8
10

V dyn,l/ Vdes,l

(c) WS Model
Fig. 4.15 Max. Dyn. Story Shears (Normalized by Vdes) for ~(SDOF) = 8, a= 0% (Mean for 15s Records)

121

0.8

.r:.

C)

2,5,10,20,30,40 stories
(thick -+ thin lines)

~ 0.6
G)

>

:; 0.4

-;

a:
0.2

0.0

0.5

1.0

1.5

2.0

2.5

3.0

V dyn,l/ V des,l

(a) BH Model
1.0

MAX. DYNAMIC STORY SHEARS, Vd n 1/ Vdes i - (15s.ch-8.1 O)

0.8

.r:.

C)

~ 0.6

2,5,1 0,20,30,40 stories


(thick -+ thin lines)

:; 0.4

-;

a:
02
0.0 +.----,-----:----~--"--'--~-....1...-_;.;;;;:.__ _---l
3.0
0.5
1~
1~
2~
2.5
0.0

Vdyn,l/ Vdes,l

(b) CH Model
1.o

MAX. DYNAMIC STORY SHEARS, V

1--T-I-r~~~~~~---1

0.8

.r:.

C)

~ 0.6
G)

>

a; o.4
-;

a:

02
o.o+----~~-~-~-~-~~~-~----~

10

v dyn,l/ vdes,l
(c) WS Model

Fig. 4.16 Max. Dyn. Story Shears (Normalized by V~)for J.tt(SDOF) = 8, a= 10% (Mean for 15s Records)

122

MAX. OYN. & DESIGN STORY SHEARS, V 11 V des base - (15s.bh40-2)


1.o r-c::;;:;:--------;:;-~~~~--iiil
- - Vdes.i IVdes.base
Vdyn.I/Vdas.base(Ct= 0%)
V .iiVdes.basa(Ct=lO%)

0.8

~
~ 0.6
Q)

>

:i 0.4
Q)

a:

0.2
0.0 +.----..,-----~--'----"~-...,....----------!
0.5
1.0
1.5
2.5
0.0
2.0

v,t vdes,base
(a) BHModel
MAX. OYN. & DESIGN STORY SHEARS, Vi IV des base- (15s.ch40-2)
1.0

- - Vdes.i I Vdes.base
- - Vdyn.i IVdes.base (Ct = 0%)
V n.;l Vdes.base (Ct = 10%)

0.8

0')

~ 0.6
Q)

>

'i

0.4

"ii

a:

0.2
o.o~----,-----~~~--,------,--------!

0.0

0.5

1.0

2.0

1.5

2.5

Vi I Vdes,base

(b) CH Model

- - VdesJ IVdes.base
- - VdynJIVdu,base(et= 0%)
V .IIVdes.base(Ct= 10%)

0.8

0')

~ 0.6
Q)_

>

'i

0.4

"ii

a:
0.2

0.0 +-----....-----~===--r---~=-----r---~
0.0
0.5
1.0
1.5
2.0
2.5

v,IVdes,base

(c) WS Model
Fig. 4.17 Max. Dyn. & Design Story Shears (Nonnalized by V~for Jlt(SOOF) 2, 40 Stories (Mean for 15s Records)

123

MAX. DYN. & DESIGN STORY SHEARS, V1/ Vdes base - (15s.bh40-8)

1"'1:!-~
. . ;:-------;;;~~~~;;;;;-ii;iil

1.0

-:!

......

- - VdesJ IVdes,base
Va,nJIVdes,bllse (a= 0%)
V ,;IVdes.bllse(a=IO%)

0.8

.c

"\.'1...

0.6

G)

>

0.4

G)

a:

0.2
0.0

.L.- - - . , - - - - . . . . : . __ _

0.0

0.5

_r:;===:_--.,---.I::=:J
1.5

1.0

2.0

2.5

3.0

Vi IV des,base
(a) BH Model
MAX. PYN. & DESIGN STORY SHEARS, v, I Vdes.base- (15s.ch40-8)
1.o ro:;;::;:--------::;;~~~~~--.,

- - Vdes,i IVdes,base
Vctvn.l/ V aes.ba.Se (a= 0%)

0.8

.C)
c

,;f V cleS.baS& (CI = JQ%)

~ 0.6
G)

>

(; 0.4
G)

a:
02
0.0 +._ _ _...,...._ _ _....__ _
0.0

0.5

~----:---.....;;;;;'i'L------1

1~

1~

2~

2.5

3.0

v,l v des,base

(b) CH Model
MAX. DYN. & DESIGN STORY SHEARS, V,l Vdes base- (15s.ws40-8)
1.0

n--c:::==~~==--~~~~~~-iiiiiil
- - Vdesj IVcles,base
Vc:~ynj/Vdes,base(a= 0%)
V ,;IVcles.basa(a=IO%)

0.8

.C)
c

:!

0.6

G)_

>

a;
a;
a:

0.4

02
0.0 -!---_..i::::==;::===~--,-----~--===----1
0
2
4
6
8

v,lvdea,base

(c) 40-Story WS Model

Fig. 4.18 Max. Dyn. & Design Story Shears (Nonnalized by Vda.b;asc) for ~(SDOF) =8, 40 Stories (Mean for 15s Records)

124

MAX. DYNAMIC STORY OVERTURNING MOMENTS, MoT(dyn,i) I MoT(des,i) {15s.20-e)


15s Records, 20-Story, Elastic, Meancr
1.0

0.8

0.6

:t:

i!
I

ii

>

t tl

0.4

..

a:

:
0.2

0.0

..

1/
!l
i/
l

'

.c

en

I
i

0.0

0.5

1.0

1.5

MoT(dyn,i) I MoT(des,i)

Fig. 4.19 Variation in Ma"<imum Dynamic Story Overturning Moments for Elastic 20-Story Structures (15s Records)

MAX. DYNAMIC STORY OVERTURNING MOMENTS, MoT(dyn,i) I MoT(des,i) (15s-e)


15s Records, Elastic, Mean
1.0

-~

'-

0.8

.c

en

'Q; 0.6

:t:
~

>

ttl

-a;

0.4

a:

0.2

0.0

~'i

~/

tv~

'/

II

0.0

0.5

2,5,1 0,20,30,40 stories


(thick~ thin lines)

1.0

1.5

MoT(dyn,i) I MoT{des,f)

Fig. 4.20 Maximum Dynamic Story Overturning Moments for Elastic Structures (Mean for 15s Records)

125

MAX. DYNAMIC STORY OVERTURNING MOMENTS, Mor

(15s.bh2.00)

1.0

-:!

0.8

.s:

_E)

0.6

G)

>

2,5,10,20,30,40 stories

0.4

(thick ~ thin lines)

C)

a:

0.2
0.0 ..1...-----~----.:..L&..------,-------..J
0.5
1.0
1.5
2.0
0.0
MoT(dyn,l) I MoT(des,l)

(a) BH Model
MAX. 0YNAMIC STORY OVERTURNING MOMENTS, Morrdvn.il/ Morrdes.il" (15s.ch-2.00)
0

0.8

.s:
C)

~ 0.6
G)

>

~ 0.4

C)

a:

0.2

2,5,10,20,30,40 storiesl
(thick -+ thin lines)

0.0 ..l..----------~-----....-----..1
1.0
1.5
2.0
0.5
0.0
MoT(dyn,l) I MoT(des,l)

(b) CHModel
MAX. DYNAMIC STORY OVERTURNING MOMENTS, M0
10

-:!

,--------::::F::c::r:c=:L..--=.;=L......;...-,

0.8

.s:::.

C)

0.6

GL

>

iii

2,5,1 0,20,30,40 stories

0.4

(thick -+ thin lines)

C)

a:
0.2
0.0

+-----r---..,.....JI....L.It..L.--..,.----.-----....------1
0.0

0.5

1.0

1.5

2.0

2.5

3.0

MoT(dyn.l) I MoT(dH,i)

(c) WS Model

Fig. 4.21 Max. Dyn. Story Ove~g Mom. (Normalized by Mandes,i)) for ~SOOF} =2, a~ 0%, (Mean for 15s Records)
126

0.8

.:::.
0

~ 0.6
C)

>

2,5, 10,20,30,40 stories


(thick-+ thin lines)

~ 0.4

a;

a:
0.2

+.----...----..a!.------.. . . .--._----1

0.0

0.0

1.0

0.5

1.5

MoT(dyn,i)

2.0

2.5

3.0

I MoT(des,l)

(a) BH Model
MA:X. ~YNAMIC STORY OVERTU~NING MOMENTS, Morcdvn.il/ Morcdes.il" (15s.ch-8.00)
1

0.8

.:::.

Cl)

~ 0.6
C)

>

~ 0.4

a;

a:

2,5,10,20,30,40 stories
(thick-+ thin lines)

0.2
0.0+---~----~---------~---~--~

0.0

1.5

1.0

0.5

2.0

2.5

3.0

I MoT(des,l)
(b) CHModel

MoT(dyn,l)

MAX. DYNAMIC STORY OVERTURNING MOMENTS, M0 d

1.0 ...-----,..---......................----~='--.:::.:..l:=L...,;.--___,

-!

0.8

.:::.
Cl)

0.6

G)

>

2,5,10,20,30,40 stories
(thick -+ thin lines)

~ 0.4

a;

a:

0.2
0.0+---LL~~~-~-----~---~-----~

MoT(dyn,l)

10

I MoT(des,l)

(c) WS Model
Fig. 4.22 ~ Dyn. Story Overturning Mom. (Nonnalized by Mor<des.iY for ~(SOOF) =8, a= 0%, (Mean for 15s Reccres;

127

0.8

.c
0

~ 0.6
Q)

>

c;
C)
a:

2,5, 10.20,30,40 stories


(thick -thin lines)

0.4
0.2

0.0

0.5

1.0

1.5

MoT(dyn,l)

2.0

2.5

3.0

I MoT(des,l)

(a) BH Model
MAX. DYNAMIC STORY OVERTURNING MOMENTS, Morrd

1.0~-------------r~~~----~~~~~~~--~

0.8

.c
0

~ 0.6
Q)

>

2,5, 10,20,30,40 stories


(thick- thin lines)

0.4

Q)

a:

0.2

0.0

1~

0.5

~0

1.5

2.5

3.0

I MoT(des,l)
(b) CH Model

MoT(dyn,l)

MAX. DYNAMIC STORY OVERTURNING MOMENTS, M011


(15s.ws-8.10)
1.0r--l-11~~~~1....-....::.:.=l.:l...,.;..-,

0.8

.c
0

~ 0.6
CD

>
t; 0.4
Ci)

a:

2,5,10,20,30.40 stories
(thick- thin lines)

0.2

0.0+-------~L-~~--~~--~~------~------~

10

MoT(dyn,J) I MoT(des,l)

(c) WS Model
Fig. 4.23 Max. Dyn. Story Overturning Mom (Norrnaliztd by Morcdes.l"V for ~(SDOF) = 8, a= 10%, (Mean for 15s Records)

128

EVALUATION OF DAMAGE POTENTIAL OF RECORDED GROUND


MOTIONS AND ITS IMPLICATION FOR DESIGN OF STRUCTURES

REPORT ON TASK 2 OF THE CUREe-KAJIMA RESEARCH PROJECT ON:

"DESIGN GUIDELINES FOR DUCTILITY AND DRIFT LIMITS"

by
Vitelmo V. Bertero
Eduardo Miranda
and
The CUREe and The Kajima Research Teams

A CUREe-KAJIMA RESEARCH REPORT


AUGUST 1991

ABSTRACT AND ACKNOWLEDGEMENTS

This report summarizes studies which have been conducted by Eduardo Miranda under the direct
supervision of Vitelmo Bertero as a part of a research project on design guidelines for ductility
and drift limits, which is carried out with the ultimate goal of developing an improved method
for earthquake-resistant design of structures. These studies, which are part of Task 2 of the
above research project, focus on the evaluation of the damage potential of recorded ground
motions and the implications of the results obtained regarding the design of structures, with
particular emphasis on the structural response modification factor for strength and the
establishment of limits for lateral drift. The studies can be considered as a complement of the
studies conducted by Helmut Krawinkler, et alia, at Stanford University and reported in
"Evaluation of Damage Potential of Recorded Ground Motions," June 1991.

The research project is supported by a grant provided by Kajima Corporation and administered
by CUREe (California Universities for Research in Earthquake Engineering). This financial
support is gratefully acknowledged. Appreciation is also expressed to the California Department
of Conservation, Division of Mines and Geology, Strong Motion Instrumentation Program
(SMIP), for providing most of the processed data of recorded ground motions used in these
studies, as well as partial support for conducting them. Several researchers of the U. S. and
Kajima project research team have provided valuable input to these studies.
discussions deserve special gratitude.

Their critical

Special thanks are expressed to Professor Helmut

Krawinkler for his helpful comments and suggestions.

TABLE OF CONTENTS

1. IN'I'R.ODUCTION . . . . . . . . . . . . . . . . .
1. 1 STATEMENT OF PROBLEM . .
1. 1. 1
Earthquake Input:
Earthquakes . . . . . . . . .
1. 2 OBJECTIVES AND SCOPE . . . .

............
............
Establishment
............
............

...
...
of
...
...

.......
.......
Reliable
.......
.......

......
......
Design
......
......

..
..

1
1

..
..

1
2

2. STATISTICAL STUDY OF INELASTIC RESPONSE SPECTRA ...........


2. 1 INTRODUCfORY REMARKS ...............................
2. 2 GROUND MOTIONS SELECfED IN THIS STUDY ...............
2. 3 INELASTIC SITE SPECTRA ................................
Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Figures ....................................................

. 5
. 5
. 5
. 6
. 11
. 15

3. STRENGTH REDUCTION SPECTRA . . . . . . . . . . . . . . . . . . . . . . . . .


3. 1 GENERAL REMARKS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3. 2 STATISTICAL STUDIES OF STRENGTH REDUCfiON FACTOR
TO DISSIPATION OF ENERGY . . . . . . . . . . . . . . . . . . . . . . .
Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. 29
. 29

.....
.....
DUE
.....
.....

. 30
. 33

4. INELASTIC DISPLACEMENT DEMAND SPECTRA ....................


4. 1 INTRODUCfORY REMARKS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4. 2 STATISTICAL STUDIES OF INELASTIC DISPLACEMENT DEMAND
SPECTRA ....................... ' . . . . . . . . . . . . . . . . . . . . .
Figures . . . . . . . . . . . ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

39
39
40
43

5. SUMMARY AND CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47


5. 1 SUMMARY ............................................. 47
5. 2 CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
6. REFERENCES ................................................. 51

1. INTRODUCTION

1. 1 STATEMENT OF PROBLEM
One of the most effective ways to mitigate the destructive effects of earthquakes is to improve
existing methods of:

designing, constructing, monitoring and maintaining new earthquake-

resistant structures (man-made facilities in general); and upgrading (retrofitting) and maintaining
existing seismic hazardous facilities. As discussed in detail in the reports on Task 1 [1] and Task
2 [2], the principal issues that remain to be solved in order to improve seismic design of new
structures and seismic upgrading of existing structures are related to the three following basic
elements: Earthquake Input to the Foundation of the Structure, Demands Imposed by This
Input on the Structure, and the Supplied Capacities to the Structure, Which Should Exceed
the Demands. Therefore, it is obvious that the essential data needed to start any reliable design
of a new structure or upgrading of an existing facility is the reliable establishment of the
earthquake input, or in other words the establishment of design earthquakes.

1. 1. 1 Earthquake Input: Establishment of Reliable Design Earthquakes. Conceptually,


the design earthquake should be that ground motion which will drive the structure to its critical
response.

In practice, the application of this simple concept meets with serious difficulties,

because, firstly, there are great uncertainties in predicting the dynamic characteristics of ground
motions that have yet to occur at the building site, and secondly, even the critical response
parameter of a specific structural system may vary according to the various limit states that could
control the design. Because in most cases the design is controlled by the safety limit state which
involves damage to the building as a result of inelastic deformations, to establish the design
earthquake at this level (limit state), it is necessary to estimate the damage potential of the
different earthquake ground motions that can occur at any given site. One of the most promising
ways, not only for estimating the damage potential of earthquake ground motions but in general
for improving the Earthquake-Resistant Design (EQRD) of structures and particularly for
improving establishment of design earthquakes for the limit states involving damage (inelastic
behavior), is through the use of an energy approach.

Reference 1 discusses in detail such an

approach based on the use of an energy balance equation.

2
As discussed in Ref. 1, and in more detail in Refs. 3 and 4, a promising parameter for improving

)
r

selection of proper design earthquakes for safety limit states is through the concept of Energy

Input, Ey , and Associated Parameters. In Refs. 3 and 4, it is shown that for defining properly
the Safety (Survival) Level design earthquake it is necessary to consider the following spectra

simultaneously: the E 1 ; the Inelastic Design Response Spectra (IDRS) for strength and
displacement; and the Energy Dissipation,

En , particularly Hysteretic Energy,

including the cumulative ductility J.L~ , and Number of Yielding Reversals (NYR) spectra.
Examples of evaluation of the E1 and IDRS (particularly for strength, i.e., the yielding strength
spectra or its equivalent yielding seismic coefficient, Cy , spectra) has been given and discussed
in Ref. 1. In Ref. 2, Professor Krawinkler and his associates address in detail not only the
problem of evaluation of the IDRS for strength, but also other important issues of seismic design
for ductility capacity, including the effects of cumulative damage.

Normalized Hysteretic

Energy Dissipation is used as the basic cumulative damage parameter. A total of 15 Western
U. S. ground motion records were considered in the statistical studies presented in Ref. 2. All
of the records are from sites corresponding to soil type S1 (rock or stiff soils). To complement
the studies conducted by Professor Krawinkler and his research associates, the authors decided
to conduct similar statistical studies of recorded ground motions, but considering also ground
motions recorded on soft soil, and obtaining not only the inelastic strength spectra but also the
inelastic deformation spectra. These studies were conducted with the following main objectives.

1. 2 OBJECTIVES AND SCOPE

The ultimate goal of the studies which are being conducted at Berkeley on the evaluation of the
damage potential of ground motions has been to improve the establishment of design earthquakes
and, consequently, to improve the earthquake-resistant design of new structures and the seismic
upgrading of existing hazardous facilities.

To achieve this goal a statistical study of 124

earthquake ground motions recorded on various soil conditions ranging from rock to very soft
soil have been evaluated, and the implications of the results obtained on the reliable establishment
of design earthquakes have been assessed [5].

3
The main objectives of this report are to summarize: first, the main results obtained in the
above studies, focussing on the normalized elastic and inelastic strength spectra, as well as on
the displacement spectra; and, second, the implications of these results regarding the
establishment of design earthquakes, particularly with reference to the use of present values for
the structural response factpr, R or Rw, and the code methodology for specifying limitation on
the interstory drift limits.

In the studies conducted, the main results of which are reported herein, the emphasis is placed
on the effects of soil, particularly on soft soils, of the seismic demands of strength and
deformation.

5
2. STATISTICAL STUDY OF INELASTIC RESPONSE SPECTRA
2. 1 INTRODUCTORY REMARKS
To improve the tools for establishing reliable design earthquakes, constant displacement ductility
ratio, Jl6

fl,

linear (p. = 1) and particularly nonlinear (p. > 1) spectra were obtained by computing

the response of a family of Single Degree of Freedom (SDOF) systems with the use of the
computer program NLSPECfRA, which was developed specifically for this purpose [5].
Constant ductility spectra were computed for 124 earthquake ground motions on various soil
conditions ranging from rock to very soft soils. The following values of ductility ratios were
selected for this study: 1 (elastic), 2, 3, 4, 5, and 6. The spectra were computed for a set of 50
periods between 0.05 and 3.0 seconds. Due to the large number of records and the computational
effort involved in calculating constant ductility nonlinear spectra, the study was limited to bilinear
systems with a post-elastic stiffness of 3% of the elastic stiffness and with a damping ratio of 5%
of critical. It has been shown that the use of elasto-plastic or bilinear hysteretic models can
provide very good estimates of strength and deformation demands of structures that show elastic
stiffness degradation under cyclic loading. Note that if the hysteretic behavior is of the elasticdeformation softening type, then the deformation demands can be considerably larger.

2. 2 GROUND MOTIONS SELECTED IN TIDS STUDY


In the last six years an extensive number of earthquake ground motions has been recorded in
different parts of the world. These ground motion records have more than doubled the number
of records previously collected.

For example, the 1987 Whittier-Narrows earthquake alone

produced more records than the total number of records obtained in the Western United States
between 1933 and 1984. For this study 124 records were selected, with emphasis on those
recorded in California and on those recorded during the last six years. Of the total number of
records, 96 (77%) were recorded in the last six years and 90 (73%) were recorded in California.

The ground motions were classified into three groups according to the soil conditions at the
recording station. These groups were rock, alluvium and very soft soil. Tables 2.1, 2.2 and 2.3
list the selected ground motions recorded on rock, alluvium and soft soil sites, respectively.

2. 3 INElASTIC SITE SPECTRA


Constant ductility nonlinear spectra were computed for all records in each soil group. Strength
demands for each record were then normalized using peak ground acceleration (PGA) and
effective peak acceleration (EPA) as defined by the ATC 3-06 recommendations [6]. Despite the
limitations of both of these quantities as scaling parameters, they were considered in this study
for the following reasons:

(1)

Despite several studies conducted in this area, there is no agreement on which parameter
should be used to scale nonlinear spectra.

(2)

Most seismic hazard studies (i.e., seismic hazard curves) and attenuation relations are
based on PGA.

(3)

Using both parameters provides an opportunity to compare the results obtained with the
results of previous studies that have used the same normalization parameters.

(4)

The results can be combined with readily available information (such as maps of peak
accelerations from maximum credible earthquakes or maps of effective peak acceleration)
that is based on these parameters.

For ground motions recorded on rock or alluvium sites, nonlinear spectra were computed for a
fixed set of 50 periods between 0.05 and 3.0 seconds. In the case of ground motions recorded
on very soft soil, spectra were computed for a fixed set of 50 ratios of Tffg , where Tg is the
predominant period of the ground. The reason for using Tffg instead ofT is that Tg can have
large variations depending on the shear wave velocity of the soil and the depth of the soft
deposits. For instance, in Mexico City, where the soft clay deposits have approximately the same
characteristics throughout the city, the predominant period of motion can vary anywhere from
0.6 seconds to more than 3.8 seconds, depending on the depth of these deposits, as well as on
the type of earthquake ground motion. For statistical analyses of spectra it makes no sense to
average spectral ordinates at a certain period for ground motions with significantly different

7
predominant periods. For structural design purposes, it is important to characterize the demands
on structures with periods shorter than, longer than, or near to the predominant period.

In this study the predominant period was computed as the period corresponding to the maximum
spectral velocity ordinate. It can be shown that essentially the same predominant period would
be obtained if the Fourier amplitude spectrum or the input energy spectrum were used instead
of the velocity spectrum, because of the relationship between these three spectra. Figure 2.1
shows six examples of how, using the velocity spectra, it is possible to identify predominant
periods, Tg, for sites in the San Francisco Bay Area, Romania, and Mexico City.

The average (mean) nonlinear strength spectra of 38 ground motions recorded at rock sites
normalized using PGA and EPA are shown in Figs. 2.2 and 2.3. The spectra are plotted for
displacement ductilities of 1, 2, 3, 4, 5 and 6 (from top to bottom). For a given ground motion,
the PGA is usually larger than the EPA, hence the spectral ordinates of the average nonlinear
spectra normalized by EPA (Fig. 2.3) are larger than those of the average nonlinear spectra
normalized by PGA (Fig. 2.2).

Mean inelastic strength demand spectra of 62 ground motions recorded on alluvium are shown
in Figs. 2.4 and 2.5, where PGA and EPA, respectively, have been used as normalizing
parameters. The maximum amplification for alluvium sites is larger than that observed at rock
sites. Mean inelastic strength demand spectra of 24 ground motions recorded on soft soil are
shown in Figs. 2.6 and 2.7.

From analysis of the plots of Figs. 2.2 to 2.7, it is clear that the spectral shapes for the inelastic
demands (J.l > 1) differ significantly from elastic (J.l

= 1) spectral shapes.

Furthermore, from

analyses of the results obtained from soft soil records it was observed that for T < Tg there is
small difference between the strength demand for ductilities between 2 and 6. This implies that
a small change in the yielding strength of a structure with T < Tg may produce a large change
in the ductility demands.

8
By comparing the average spectra of the three different soil conditions it can be seen that the
largest dynamic amplification for elastic response (fl

= 1) is produced for soft soil sites.

These

results are different .:rom those reported in Ref. 7, where larger amplifications for rock and
alluvium sites than for soft soil sites are shown. Moreover, the maximum amplification (with
respect to the PGA) computed in that reference is nearly 30% smaller than the maximum
amplification computed here. For rock and alluvium sites, the maximum amplifications computed
in this study are practically the same as those reported in Ref. 7, with a smaller set of ground
motions.
Regardless of the type of local site conditions, for ductility demands equal to or larger than 5
strength demands decrease monotonically with increasing period.

While mean spectra provide information on the most probable demands of the structure, it is
important to take into account the dispersion of these demands. One method of measuring the
dispersion is to compute the coefficient of variation (COV), which is defined as the ratio of the
standard deviation to the mean. Figures 2.8 to 2.13 show the COV of the spectral ordinates for
the three types of soil conditions, with PGA and EPA used as normalizing parameters. The use
of acceleration parameters to normalize the spectra produces an increase in dispersion with
increase in period. Relatively large COVs are produced in the long period region; however,
strength demands are usually very small in this region. For soft soil sites, a local increase in
COV is produced for periods near the predominant period (i.e., T{f8

=1).

The use of peak

ground displacement as the normalizing parameter would produce an opposite trend (i.e., small
COVs in the long period range and large COVs in the short period range). By comparing COVs
of spectra normalized by PGA and EPA, it can be seen that except for periods between 0.1 and
0.5 second, COVs are larger for spectra normalized using EPA than for spectra normalized using
PGA. An important observation is that, regardless of the type of soil, COVs are nearly the same
for different levels of ductility, which means that dispersion in strength demands does not
increase with increasing ductility demands.

Mean plus one standard deviation inelastic strength demands for the three types of soil conditions

.l

9
are shown in Figs. 2.14 to 2.19.

Spectral amplifications between 0.1 and 0.5 second for ground motions recorded on soft soil are
usually much smaller than the 2.5 factor which is used for defining EPA. Therefore, the use of
EPA (which was primarily developed for firm soil records) to normalize strength demands on soft
soil sites produces unusually large amplifications (see Figs. 2.7 and 2.19) For this type of soil
condition, the use of PGA is probably more appropriate than the use of EPA.

11
STATION NAME

GEOLOGY

EAR11-lQUAKE
DATE

SAN FRANCISCO

SIDceous

San Francisco

Golden Gate Pari<

sandstone

March 22, 1957

PARKFIELD
Cholame Shandon No.2
CASTAIC
Old Ridge Road
LLOLLEO

VALPARAISO

LA UNION

LAVIWTA

ZIHUATANEJO

Rock

Sandstone

Pari<filed
June 27, 1966
San Fernando
February 9, 1971

Sandstone &

Central Chile

volcanic rock

March 3, 1985

Volcanic

Central Chile

rock

March 3, 1985

Metavolcanic

Michoacan

Rock

Sept. 19, 1985

Gabbro

Michoacan

Rock

SGpt. 19, 1985

Tuna lite

Michoacan

Rock

SGpt 19, 1985

NATL GEOGR.

Balsamo

San Salvador

INSTITUTE

Formation

October 10,1986

INST. URBAN

Ruviate

San Salvador

CONSTRUCTION

Pumice rock

October 10,1 986

GEOTECH. INVEST.

Ruviate

San Salvador

CENTER

Pumice rock

October 10,1986

MTWILSON

Quartz

Whittier-Narrows

Caltech Seismic Station

dionte

October 1, 1987

CORRAUTOS

Landslide

Lema Prieta

E:.zreka Canyon Road

deposits

Oc'.ober 17, 1989

SANTA CRUZ

Lema Prieta

ucsc

Umestone

SAN FRANCISCO

Franciscan

Lema Prieta

Cliff House

sandstone

October 17, 1989

SAN FRANC!SCO

Franciscan

Lema Prieta

Pacific Heights

sandstone

October 17, 1989

SAN FRANCISCO
Presidio

Serpentine

Oc'.ober 17, 1989

Lorna Prieta
October 17, 1989

SAN FRANCISCO

Franciscan

Rincon Hill

sandstone

October 17, 1989

YERBA BUENA

Franciscan

Lema Prieta

ISLAND

sandstone

Lema Prieta

October 17, 1989

I MAGN.

EPICTR.
DlST.[km]

5.3(MJ

11

5.6(MJ

6.5(MJ

29

7.8(Ms)

45

7.8(Ms)

84

8.1(Ms)

84

8.1 (Ms)

44

8.1 (Ms)

135

5.4(Ms)

5.7

5.4(Ms)

5.3

5.4(Ms)

4.3

6.1(MJ

19

7.1(Ms)

7.1 (Ms)

16

7.1(Ms)

99

7.1(Ms)

97

7.1(Ms)

98

7.1(Ms)

95

7.1(Ms)

95

PGA

PGV

[g's]

[c:nlsec]

N10E

0.08

4.91

S80E

0.11

4.61

N55E

0.48

78.09

N21E

0.32

17.16

N69W

0.27

27.82

DIRECTION

N10E

0.67

23.70

S80E

0.43

43.60

N70E

0.18

16.30

S20E

0.16

9.70

NOOE

0.17

20.34

N90E

0.15

11.70

NOOE

0.13

i 6.11

N90E

0.12

10.51

N90W

0.10

15.86

SOOE

0.16

18.34

270

0.53

72.70

180

0.39

56.10

90

0.38

39.20

180

0.67

55.50

180

0.42

61.80

90

0.68

80.00

90
. 360

0.19

4.09

0.13

4.32

90

0.47

47.50

360

0.62

55.20

90

0.41

21.20

360

0.43

21.20

90

0.11

21.00

0.07

11.20

360

0.05

9.88

270

0.06

14.30

90

0.20

33.50

0.10

13.30

90

0.09

11.50

360

0.08

7.34

90

0.06

14.70

360

0.03

4.61

Table 2.1 Selected ground motions recorded at rock sites

12

STATION NAME

ELCENTRO
ln1gallon District
TAFT
Uncoln School Tunnel
FIGUEROA
445 Figueroa St.
HOLLYWOOD
FrEH~

Field

Alh.Nium

Alwium

AJII.Nium

Alii.Nium

EARTHQUAKE
DATE
Imperial Valley
May 18, 1940
Kern County
July 21' 1952

February 9, 1971
San Fernando
February 9, 1971

AVE.STARS

Silt & sand

San Fernando

layers

February 9, 1971

Kokutetsu Bldg.
MELOLAND
Interstate 8 Overpass
BONDS CORNER
Highways 98 & 115
JAMES ROAD
El Centro Array 11 5

IMPERIAL V. COLLEGE
El Centro Array il 7
ELALMENDRAL

VINA DEL MAR

ZACATULA

Alii.Nium

Alluvium

Alluvium

Allwium

AJiwium

BURBANK
Cal. Fed. Savings Bldg.

Imperial Valley
October 15, 1979
Imperial Valley
October 15, 1979
Imperial Valley
October 15, 1979
Imperial Valley
October 15, 1979

fill

March 3, 1985

Alh.Nial

Central Chile

sand

March 3, 1985

Allwium

ALTADENA
Eaton Canyon Park

June 12, 1978

Central Chile

ALHAMBRA
Freemont School

MlyagiKe~Oki

Compacted

Allwium

Alluvium

Alluvium

MAGN.

s.3<Mu

7.7(Ms)

San Fernando

1901 Ave. of the Stars


SENDAI CITY

GEOLOGY

Mlchoacan
Sect. 19,1985
Whittier-Narrows
October 1, 1987

E?ICTR.
DIST.[km]

a
56

s.s<Mu

41

s.s<Mu

35

s.s<Mu

38

7.4(Ms)

110

s.s(Mu

21

6.6(MJ

6.6(MJ

22

6.s(Mu

21

7.B(Ms)

84

7.8(Ms)

8.1 (Ms)

sa.
49

6.1(MJ

s.1(Mu

13

6.1(MJ

26

Whittier-Narrows
October 1, 1967
Whittier-Narrows
October 1, 1987

PGA

PGV

[g's]

[em/see]

S90W

0.21

36.92

SOOE

0.34

33.45

DIRECTION

N21E

0.15

15.72

SSSE

0.17

17.71

N52!:

0.15

17.38

S38W

0.12

17.31

N90E

0.21

21.14

scow

0.17

16.50

N4SN

0.14

9.65

S44W

0.15

16.74

N90W

0.44

57.01

NOOE

0.24

36.16

360

0.31

71.65

270

0.30

90.94

S40E

0.58

43.63

S50W

0.77

44.07

S40E

0.52

43.99

ssow

0.37

86.56

S40E

0.33

44.56

S5ow

0.45

107.80

N50E

0.29

26.90

S40E

0.16

19.80

N70W

0.23

27.70

S20W

0.36

33.20

SOOE

0.26

30.39

N90W

0.18

13.96

270

0.40

16.98

180

0.30

21.63

90

0.16

4.92

360

0.31

10.43

130

0.22

12.61

40

0.17

9.63

Table 2.2 Selected ground motions recorded at alluvium sites

'

13

STATION NAME

GEOLOGY

EARTHQUAKE
DATE

DOWNEY

County Maint Bldg.

Deep alhN!um

Whittier-Narrows
October 1, 1987

INGLEWOOD

Terrace

Whittier-Narrows

Union 011 Yard

deposits

October 1, 1987

LOS ANGELES

Terrace

Whittier-Narrows

116th Sl School

deposits

October 1, 1987

LOS ANGELES

Allt.Mum

Whittier-Narrows

Baldwin Hills

over shale

October 1, 1987

LOS ANGELES
Hollywood Storage FF
LOS ANGELES
Obregon Parle
LONG BEACH
Rancho Los Cemtos
SAN MARINO
Southwestern Academy
TARZANA
Cedar Hill Nursery
WHITTIER
7215 Blight Tower
ALBA
900 S. Fremont
CAPITOLA
Fire Stabon
HOLLISTER
South & Pine
OAKLAND
2-StOI"f Office Bldg.
STANFORD
Parking Garage

Allt.Mum

Alluvium

Allt.Mum

Alluvium

Alluvium

Alluvium

Alluvium

Allt.Mum

Alluvium

Alluvium

Allt.Mum

Whittier-Narrows
October 1, 19 87
Whittier-Narrows
October 1, 1987
Whittier-Narrows
October 1, 19 87
Whittier-Narrows
October 1, 1987
Whittier-Narrows
October 1, 1987
Whittier-Narrows
October 1, 1987
Whittier-Narrows
October 1, 1987
Lema Prieta
October 17, 1989
Lema Prieta
October 17, 1989
Lema Prieta
October 17, 1989
Lema Prieta
October 17, 1989

MAGN.

EPICTR.
DIST.(km]

6.1(MJ

17

6.1(MJ

25

6.1(MJ

22

6.1(MJ

27

6.1(MJ

25

6.1(MJ

10

6.1(MJ

27

6.1(MJ

6.1(MJ

44

6.1(MJ

10

6.1(MJ

7.1 (Ms)

7.1 (Ms)

48

7.1(Ms)

92

7.1 (Ms)

51

PGA

PGV

(g's]

[em/sec]

270

0.16

12.76

180

0.20

29.27

DIRECTION

90

0.23

16.28

360

0.27

8.65

360

0.40

18.65

270

0.29

17.83

90

0.17

6.64

360

0.15

7.51

90

0.12

6.94

360

0.21

8.99

360

0.44

22.07

270

0.45

12.98

90

0.25

18.48

360

0.15

~6.65

360

0.20

12.87

270

0.15

4.82

90

0.63

24.20

360

0.46

19.27

90

0.63

27.10

360

0.43

26.60

90

0.29

10.84

360

0.25

20.64

90

0.39

30.70

360

0.46

36.10

90

0.17

30.90

360

0.36

62.80

290

0.24

37.90

200

0.19

20.00

360

0.26

33.18

90

0.22

21.30

Table 2.2 Selected.ground motions recorded at alluvium sites (cont.)

14

STATION NAME

GC:OLOGY

EARTHQUAKE

MAGN.

DATE

BUCHAREST
Building Research lnst.
SCT

Stia. de

Soft

Soft

Romania
March 4 ,19n

clay

Sept 19, 1985

CENTRAL DE ABASTOS

Soft

Michoac~n

Frigor1fico

clay

CENTRAL DE ABASTOS

Soft

Mlchoa~n

Ofidna

clay

Sept19, 1985

COLONIA ROMA

EMERYVILLE
Free Field South

Soft

Acapulco

clay

April 25, 1989

Say mud

Say mud

OAKLAND
Outer Harbor Wharl

Say mud

Fill

SAN FRANCISCO
International Airport

October 17, 1989

Say mud

October 17, 1989


Fill over

Lorna Prieta

18-Story Comerdal Sldg.

bay mud

October 17, 1989

FOSTER CITY
Say mud

174

8.1(Ms)

8.1(Ms)

8.1 (Ms)

6.9(Ms)

385

389

389

7.1(Msl

97

7.1 (Ms)

97

7.1 (Ms)

95

7.1 (Ms)

98

Lorna Prieta

SAN FRANCISCO

Redwood Shores

7.1(Ms)

Lorna Prieta
October 17, 1989

Lorna Prieta
October 17, 1989

(cmlsacj

'2N

0.17

32.62

SN

0.20

75.11

N9CJN

0.17

60.50

SOOE

0.10

38.74

99.53

0.10

34.57

n.52

0.08

24.85

76.50

0.08

41.86

67.95

0.07

34.98

N90W

0.06

11.90

SOOE

0.05

10.92

350

0.21

21.50

260

0.26

41.06

Lorna Prieta
October 17, 1989

TREASURE ISLAND
Naval Base

Lorna Prieta

(g's]

DIST.[km]

Lorna Prieta
October 17, 1989

EMERYVILLE
Free Field North

Sept 19, 1985

PGV

DIRECTION

Michoac~n

Comunic. y Transport.

PGA

EPICTR.

7.1(Ms)

79

7.1(Ms)

$5

7.1(Ms)

63

350

0.20

15.74

260

0.22

37.94

305

0.27

42.30

125

0.29

40.80

90

0.15

33.40

360

0.10

15.60

90

0.33

29.25

350

0.23

26.45

980

0.13

17.11

350

0.15

15.76

90

0.28

45.40

0.25

31.80

Table 2.3 Selected ground motions recorded at soft soil sites

15

VELOCITY (in/sec)

VELOCI1Y On/sec)
~0

25

FOSTER CITY 0

SF 18 STORY BLDG. 3SO

T;- t.tS

40

20

30

15

20

10

10

0
0.0

1.0

0.5

1.5

2.0

2.5

3.0

0.0

0.5

PERIOD (sec)

1.0

1.5

2.0

2.5

3.0

2.5

3.0

PERIOD (sec)

VELOCITY (in/sec)

VELOCI1Y (in/sec)

60

70

EMERYVlLLE Fi=S 260

60

50

50

40

40

30
30
20

20

10

10

0.0

1.0

0.5

1.5

2.0

2.5

3.0

0.0

0.5

1.0

1.5

2.0

PERIOD (sec)

PERIOD (se:c)

VELOCITY (In/sec)

VELOCI1Y On/sec)

140

25

SCT EW

COLONIA ROMA N90W


1.2s

Tr

120

20

100
15

eo
60

10

40
5

20
0

0
0.0

0.5

1.0

1.5

2.0

PERIOD (sec)

2.5

3.0

0.0.

0.5

1.0

1.5

2.0

PERIOD (sec)

Figure 2.1 Predominant ground period for various soft soil sites.

2.5

3.0

16

11
3.0
ROCK SITES

2.5

= 1

2.0
1.5
1.0

0.5
0.0
0.0

0.5

1.0

1.5
2.0
PERIOD (sec)

2.5

3.0

Figure 2.2 Mean strength demands of ground motions recorded on rock when normalized using PGA (!-1=1 ,2,3,4,5,6}.

Cy
EPNg

3.0
ROCK SITES

2.5
2.0
1.5
1.0

0.5
0.0
0.0

0.5

1.0

1.5
2.0
PERIOD (sec)

2.5

3.0

Figure 2.3 Mean strength demands of ground motions recorded on rock when normalized using EPA (J.1=1 ,2,3,4,5,6).

17

'1
3.0

ALLUVIUM SITES

2.5

= 1

2.0
1.5
1.0

0.5
0.0
0.0

Figure 2.4

0.5

1.0

1.5
2.0
PERIOD (sec)

2.5

3.0

Mean strength demands of ground motions recorded on alluvium when


normalized using PGA (J..L=1 ,2,3,4,5,6).

Cy

EPNg

3.0
ALLUVIUM SITES

2.5
2.0
1.5
1.0

0.5
0.0
0.0

0.5

1.0

1.5
2.0
PERIOD (sec)

2.5

3.0

Figure 2.5 Mean strength demands of ground motion recorded on alluvium when normalized using EPA (!-1=1 ,2,3,4,5,6).

18

11
5

SOFT SOIL SITES

2
1

0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

T /Tg
Figure 2.6 Mean strength demands of ground motions recorded on soft soil when
normalized using PGA (!l=1,2,3,4,5,6).

Cy
EPNg
5
SOFT SOIL SITES

0.0

0.5

1.0

1.5

2.0

2.5

3.0

T /Tg
Figure 2.7

Mean strength demands of ground motions recorded on soft soil when


normalized using EPA (!l=1,2,3,4,5,6).

19
cov

cov

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4
0.2

0.2

J.1=2

!l = 1
0.0

0.0
0.0

0.5

1.0
1.5
2.0
PERIOD (sec)

2.5

3.0

0.0

cov

cov

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.5

1.5
2.0
1.0
PERIOD (sec)

2.5

3.0

0.2

IJ.=4

J.1=3
0.0

0.0
0.0

0.5

1.0
1.5
2.0
PERIOD (sec)

2.5

3.0

0.0

cov

cov

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.5

1.0
1.5
2.0
PERIOD (sec)

2.5

3.0

0.2

1J.=6

IJ.=S
0.0

0.0
0.0

0.5

Figure 2.8

1.0
1.5
2.0
PERIOD (sec)

2.5

3.0

0.0

0.5

2.0
1.5
1.0
PERIOD (sec)

2.5

3.0

Coefficients of variation of strength demands (normalized by PGA) for


ground motions recorded on rock.

20

cov

cov

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4
0.2

0.2
j.L

=1

j.L=2
0.0

0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0

0.5

PERIOD (sec)

1.0

1.5

2.0

2.5

PERIOD (sec)

cov

cov

1.0

1.0

IJ.8

0.8

0.6

0.6

0.4

0.4

0.2

3.0

0.2

j.L=3

j.L=4

0.0

0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0

0.5

1.0

1.5

2.0

2.5

3.0

PERIOD (sec)

PERIOD (sec)

cov

cov

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

j.L=5

j.L=6

0.0

0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0

PERIOD (sec)

0.5

1.0

1.5

2.0

2.5

3.0

PERIOD (sec)

Figure 2.9 Coefficients of variation of strength demands (normalized by EPA) for


ground motions recorded on rock.

21
cov

cov

1.2

1.2

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

!.1 = 1

J..l=2

0.0

0.0
0.0

0.5

1.0

1.5

2.0

2.5

0.0

3.0

0.5

cov

cov

1.2

1.2

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

1.0

1.5

2.0

2.5

3.0

PERIOD (sec)

PERIOD (sec)

0.2

J..l=3

J..l=4

0.0

0.0
0.0

0.5

1.0

1.5

2.0

2.5

0.0

3.0

0.5

cov

cov

1.2

1.2

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

1.0

1.5

2.0

2.5

3.0

2.5

3.0

PERIOD (sec)

PERIOD (sec)

0.2

J..l=S

0.0

0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0

PERIOD (sec)

0.5

1.0

1.5

2.0

PERIOD {sec)

Figure 2.10 Coefficients of variation of strength demands (normalized by PGA) for


ground motions recorded on alluvium.

22

cov

cov

, .2

1.2

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

1.1=1

1.1=2

0.0

0.0
0.0

0.5

1.0

1.5

2.0

2.5

0.0

3.0

0.5

PERIOD (sec)

1.5

2.0

2.5

3.0

2.5

3.0

2.5

3.0

PERIOD (sec)

COV

cov

1.2

1.2

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

1.0

0.2

1.1=3

0.0

0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0

0.5

PERIOD (sec)

1.0

1.5

2.0

PERIOD (sec)

cov

cov

1.2

1.2

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

1.1=5

0.0

0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0

PERIOD (sec)

0.5

1.0

1.5

2.0

PERIOD (sec)

Figure 2.11 Coefficients of variation of strength demands (normalized by EPA) for


ground motions recorded on alluvium.

23

cov

cov

1.2

1.2

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

!..l = 1

!..l=2
0.0

0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0

0.5

T /Tg

1.0

1.5

2.0

2.5

T /Tg

cov

cov

1.2

1.2

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

!..l=3

0.0

3.0

!..l=4

0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0

0.5

1.0

T /Tg

1.5.

2.0

2.5

3.0

T /Tg

cov

cov

1.2

1.2

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

0.0

!..l=6

0.0
00

0.5

1.0

1.5

T /Tg

2.0

2.5

3.0

0.0

0.5

1.0

1.5

2.0

2.5

3.0

T /Tg

Figure 2.12 Coefficient of variations of strength demands (normalized by PGA) for


ground motions recorded on soft soil.

24

cov

cov

i.2

1.2

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

1.1=2

0.0

0.0
0.0

0.5

1.0

1.5

T/Tg

2.0

2.5

0.0

3.0

cov

cov

1.2

1.2

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.5

1.0

1.5

T /Tg

2.0

0.2

1.1=3

0.0

2.5

3.0

!J.=4

0.0
0.0

0.5

1.0

1.5

2;0

T/Tg

2.5

0.0

3.0

cov

0.5

1.0

1.5

2.0

T /Tg

2.5

3.0

cov

1.2

1.2

1.0

1.0

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

!J.=S
0.0

!J.=6

0.0
0.0

0.5

1.0

1.5

2..0

2.5

3.0

0.0

T /Tg

0.5

1.0

1.5

2.0

2.5

3.0

T /Tg

Figure 2.13 Coefficients of variation of strength demands (normalized by EPA) for


ground motions recorded on soft soil.

25

ROCK SITES

=1

0
0.5 .

0.0

1.0

1.5
2.0
PERIOD (sec)

2.5

3.0

Figure 2.14 Mean plus one standard deviation strength demands (normalized by PGA)
for rock sites.

Cy

EPNg
4

!J.

=1

ROCK SITES

!J.

0.0

0.5

1.0

1.5
2.0
PERIOD (sec)

2.5

3.0

Figure 2.15 Mean plus one standard deviation strength demands (normalized by EPA)
for rock sites.

26

11
4
ALLUVIUM SITES
I.A.

= 1

0
0.0

0.5

1.0

1.5
2.0
PERIOD (sec)

2.5

3.0

Figure 2.16 Mean plus one standard deviation strength demands (normalized by PGA)
for allwium sites.

cy
EPNg
4
ALLWIUM SITES

0.0

0.5

1.0

1.5
2.0
PERIOD (sec)

2.5

3.0

Figure 2.17 Mean plus one standard deviation strength demands (normalized by EPA)
for allwium sites.

27

i1
8
SOFT SOIL SITES

2
J.l.

0
0.0

=6
0.5

1.0

1.5

2.0

2.5

3.0

T /Tg
Figure 2.18 Mean plus one standard deviation strength demands (normalized by PGA)
for soft soil sites.

Cy

EPNg
8
SOFT SOIL SITES

2
J.l.

0.0

=6
0.5

1.0

1.5

2.0

2.5

3.0

T /Tg
Figure 2.19 Mean plus one standard deviation strength demands (normalized by EPA)
for soft soil sites.

29
3. STRENGTH REDUCTION SPECTRA
3. 1 GENERAL REMARKS
Current earthquake-resistant design philosophy accepts structural and non-structural damage in
the event of severe earthquake ground motions. A consequence of this philosophy is that code
design forces are smaller, and in some cases much smaller, than those required to maintain elastic
behavior.

As discussed in Ref. 1, in present U.S. seismic codes the reduction in forces is done through the
use of structural modification factors (R or Rw). In the commentaries on these codes it is stated
that these R and Rw factors should account for at least dissipation of energy through plastic
behavior (p.), increase in damping, and the inherent overstrength in the structure. Nevertheless,
there is no rational procedure from which the recommended values have been derived, and, as
presently specified, they are empirical and only a function of the structural system, thus assuming
that the R or Rw are the same regardless of the values of the period of vibration, T 1 , or the local
site conditions. Moreover, the conglomeration of so many variables into one factor obscures the
underlying intent, and greatly impedes the development of improved understanding.
Improvements in the design process can be achieved if the factors contributing to the reduction
of seismic forces are taken into account explicitly. As discussed in Ref. 1, the R factor can be
defined as

(3. 1)
where:

R~

the reduction due to dissipation of energy through hysteretic behavior.

R;

the reduction due to change in damping ratio from elastic to inelastic


behavior.

the reduction due to the real strength of the constructed structure at the
time that the earthquake occurs compared with the code required
design strength: Rs

= Rovs

+1

Rovs the reduction due to the overstrength between the real strength and the
code required strength.

30
The study reported herein was devoted to obtaining reliable data on the quantification of R11

paying particular attention to how the values of R11 are affected by local soil conditions, including
very soft soil.

3. 2

STATISTICAL STUDIES OF STRENGTH REDUCTION FACTOR DUE TO

DISSIPATION OF ENERGY
In this study an attempt has been made to estimate the reduction in strength that is produced by
allowing inelastic behavior.

This reduction in strength demand due to nonlinear behavior

(referred to in this study as R) is given by the ratio of the elastic demand to the strength demand
on an inelastic system undergoing a certain ductility demand, 1Ji ,

R"

= C1 (~=1)

(3. 2)

Cy(~=~i)

This ratio was obtained for a total of 31,000 different SDOF systems (the product of 124 ground
motions, 50 periods and 5 levels of displacement ductility). Figure 3.1 shows the mean of the
strength demand reductions due to nonlinear behavior for ground motions recorded on rock. It
can be seen that, unlike code-specified reduction factors, force reductions are period dependent.
In general, the reductions are small in the short-period range, increase to values larger than p, for
periods between 0.8 and 2.5 seconds, and approach p, for periods greater than 2.5 seconds.
Maximum reductions are produced at a period of approximately 1.3 seconds.

Reductions

corresponding to mean minus one standard deviation are shown in Fig. 3.2. Strength reduction
factors for alluvium and soft soil sites are shown in Figs. 3.3 and 3.4. Strength reductions for
alluvium sites follow, in general, the same trend as strength reductions for rock sites. For periods
shorter than 1.2 seconds, reductions for large p, (p. > 2) computed for alluvium sites are larger
than those computed for rock sites.

For soft soils and periods close to the predominant period of the site, very large reductions are
produced. For the same displacement ductility ratio (larger than 4) these reductions are nearly
twice the maximum reductions computed for rock or alluvium sites. For periods greater than two
times the predominant site period (T!fg > 2.0) strength reductions are approximately equal top,.

31
ForT< 0.9Ts, the reductions are small, smaller than fl. The sharp decrease of

R~

with T < Ts

compared toT> Tg points out clearly the importance of estimating as accurately as possible the
Ts and the T of the structure. Because of the uncertainties involved in such estimations, it will
be necessary to consider a band of values of Tg as well as ofT.

In order to study the dispersion of the computed strength reduction factors, coefficients of
variation (COV) were computed for each type of soil condition and each different displacement
ductility ratio. Dispersion of computed values of R.u are shown in Figs. 3.7, 3.8 and 3.9, where
COVs are presented for each type of soil condition and four levels of displacement ductility.
Unlike normalized strength demands, COVs of R.u do not show important variations with changes
in period, and increase slightly with increasing displacement ductility.

33

RJ.L
8
J.L=6

ROCK SITES

0
0.0

0.5

1.0

1.5
2.0
PERIOD (sec)

2.5

3.0

Figure 3.1 Mean of strength reductions due to nonlinear behavior for rock sites.

R!!
8
ROCK SITES

0.0

Figure 3.2

0.5

1.0

1.5
2.0
PERIOD (sec)

2.5

3.0

Mean minus one standard deviation of strength reductions due to nonlinear behavior for rock sites.

34
R~
8
ALLUVIUM SITES

Jl=6

!.1=5
Jl=4

!.1=3

!J.=2

0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

PERIOD (sec)

Figure 3.3 Mean of strength reductions due to nonlinear behavior tor alluvium sites.

R~
8
ALLWIUM SITES

0.0

Figure 3.4

0.5

1.0

1.5
2.0
PERIOD (sec)

2.5

3.0

Mean minus one standard deviation of strength reductions due to nonlinear behavior for alluvium srtes.

35

RJ.L
i4

SOFi SOIL SITES

'12
'10

8
6

0.0

0.5

1.0

1.5

2.0

2.5

3.0

T /Tg
Figure 3.5

Mean of strength reductions due to nonlinear behavior for soft soil sites.

SOFi SOIL SITES

'12
10

1-4

6
4

2
0

0.0

0.5

1.0

1.5

2.0

2.5

3.0

T /Tg
Figure 3.6

Mean minus one standard deviation of strength reductions due to nonlinear behavior for soft soil sites.

36

cov

cov
, .0

1.0
!-l=2

!-l=3

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

0.0

0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0

0.5

PERIOD (sec)

1.0

1.5

2.0

2.5

3.0

PERIOD (sec)

cov

cov

1.0

1.0
!-l=4

!-l=5

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

0.0

0.0
0.0

0.5

1.0

1.5

2.0

PERIOD (sec)

Figure 3.7

2.5

3.0

0.0

0.5

1.0

1.5

2.0

2.5

3.0

PERIOD (sec)

Coefficients of variation of strength reductions due to nonlinear behavior


tor rock sites.

37

cov

cov

1.0

1.0

!-1=3

!J.=2
0.8

0.8

0.6

0.6

0.4

0.4

0.2

~~

0.0

0.2

0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0

0.5

PERIOD (sec)

1.0

1.5

2.0

2.5

3.0

PERIOD (sec)

cov

cov

1.0

1.0

!-1=5

!J.=4
0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

0.0

0.0
0.0

0.5

1.0

1.5

2.0

PERIOD (sec)

2.5

3.0

0.0

0.5

1.0

1.5

2.0

2.5

3.0

PERIOD (sec)

Figure 3.8 Coefficients of variation of strength reductions due to nonlinear behavior


for allwium sites.

38

cov

cov

1.0

1.0
J..L=2

J..L=3

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

0.0

0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0

0.5

1.0

T /Tg

1.5

2.0

2.5

3.0

T/Tg

cov

cov

1.0

1.0
J..L=4

J..L=5

0.8

0.8

0.6

0.6

0.4

0.4

0.2

0.2

0.0

0.0
0.0

0.5

1.0

1.5

2.0

T /Tg

2.5

3.0

0.0

0.5

1.0

1.5

2.0

2.5

3.0

T/Tg

Figure 3.9 Coefficients of variation of strength reductions due to nonlinear behavior


for soft soil sites.

39
4. INELASTIC DISPLACEMENT DEMAND SPECTRA

4. 1 INTRODUCTORY REMARKS
In Ref. 1, the importance and the need for controlling the relative displacements (drift) of a
structure when subjected to seismic excitations is discussed in detail. This reference contains the
following statements:
"While displacement ductility factors generally provide a good indication of structural
damage, they do not usually adequately reflect the damage to nonstructural elements.
This is an important limitation in seismic-resistant design since a significant portion of
the hazard to occupants and of the total cost of repairing earthquake damage is a
consequence of nonstructural damage. Nonstructural damage is more dependent on the
relative displacements (drift) than on the overall displacements. To obtain a reliable
measure of nonstructural damage, maximum drifts must remain unnormalized or be
divided by the value of drift corresponding to the damage threshold.

Nonstructural

damage estimates based on drift ductilities may be misleading.

For example,

nonstructural damage for relatively rigid structures may be small even for large values of
displacement ductility since the yield displacement may be well below the nonstructural
damage threshold. On the other hand, the nonstructural damage and lateral displacements
for flexible structures may become intolerable large even before significant yielding
develops.

To produce safe and economical structures, seismic resistant design methods must
incorporate drift (damage) control in addition to lateral displacement ductility as design
constraints."

"The growing concern over the costs of earthquake damage (direct, functional and
indirect) and the difficulty of repairing much postyield damage, points out the need that
more attention should be given to control of damage and repairability at the design stage.
These needs have been clearly emphasized by the 1989 Lorna Prieta earthquake. The
control of damage will, of course, also help to improve human safety, which is the

40

traditional fundamental criterion. The main source of damage is deformation; thus, to


control damage, it is necessary to control deformation and particularly to control interstory
drift.

In summary, achievement of reliable and efficient EQRDs requires satisfaction not only
of the criterion for strength and toughness, but also the criteria for deformation and
repairability.

It should be noted that strength, toughness, deflection control and

repairability are interrelated and hard to define."

As discussed in Ref. 1, it is believed that one of the most practical methods for including the

criterion of IDI control in the preliminary design is to use displacement spectra for the different
levels of earthquakes that are included in the design. For the case of two levels of design
earthquakes, service and safety (survival) levels, a procedure has been suggested in Ref. 8. The
method is based on the use of Smoothed Design Response Spectra (SDRS) in tripartite
logarithmic scale, as discussed in Ref. 9. For the serviceability limit state, the displacements can
be read directly from the SLEDRS for strength. For the safety limit state, although there is a
need for two spectra (one for the strength and one for the displacement), the IDRS for
displacement can be obtained directly from the IDRS for strength for any given ductility ratio,
p., by multiplying their ordinates by the corresponding p.. This approach is discussed below.

From the point of view of present practice, it was considered of interest to try to find out to what
extent inelastic displacement demands can be predicted using elastic analysis. Thus, the ratio of
inelastic to elastic deformation was computed for each time-history analysis, and for a set of 50
periods between 0.05 and 3.0 seconds.

4. 2 STATISTICAL STUDIES OF INElASTIC DISPlACEMENT DEMAND SPECTRA


Figure 4.1 shows mean displacement ratios (~elastic l~tastic) for ground motions recorded on rock.
As observed in previous studies [9], in the short-period range inelastic displacement demands can

be considerably larger than elastic demands. Previous studies have recommended fixed periods
(independent of p.) to specify spectral regions in which elastic analysis can be used to estimate

41
inelastic displacement demands. However, it can be seen that this period is clearly dependent
on the ductility level.

For instance, for a ductility ratio of 2, inelastic displacements are

approximately the same as elastic displacements for periods greater than 0.5 second, while for
a ductility ratio of 6 such an assumption is only valid for periods greater than about 1.0 second.

Using the normalized strength demands in Figs. 2.2 to 2.7, elastic and inelastic displacement
demands can be obtained using the expression

(4. 1)
or

= (CjEPA)I-LT (EPA)

(4. 2)

41t2

Figure 4.2 is an example of a displacement demand spectra which has been computed assuming
an EPA of 0.4g, consistent with that assumed by ATC 3-06 [6] for the regions of highest seismic
risk in the U.S. Even though inelastic displacement demands can be considerable larger than
elastic demands for very short periods (T < 0.2 second), displacement demands in this region are
very small.

Displacement ratios for ground motions recorded on alluvium are shown in Fig. 4.4.
Displacement demands assuming an EPA of 0.4g are shown in Fig. 4.4. In this case, for a
ductility ratio of 2 inelastic displacements are approximately the same as elastic displacements
for periods greater than 0.3 second.

While Figs. 4.2 and 4.4 give the most probable displacement demands for rock and alluvium sites
(because they are based on mean values), caution should be exercised in the use of these spectra
or any other mean displacement spectra which have been computed from spectra normalized by
acceleration parameters (i.e., PGA or EPA) which result in large dispersion in the long-period

42
range.

If more reliable displacement demands are desired for long-period structures, it is

recommended that the inelastic strength demand spectra be normalized by velocity or


displacement parameters.

Mean ratios of inelastic to elastic displacements for ground motion records on soft soil are shown
on Fig. 4.5. For periods near the predominant period of the site, inelastic displacements are 30%
to 40% smaller than the corresponding elastic displacements. This observation has important
design implications. For instance, the use of energy dissipation devices can be particularly
effective in this situation, where not only very large force reductions may be achieved, but also
(and more importantly for damage control) significant reductions may be obtained in
displacement demands. However, these reductions should be used cautiously for design, because
they depend on accurate estimations of T and Tg.

For periods shorter than the predominant site period, inelastic displacement demands can be
significantly larger than elastic demands, which for sites with very long periods, T g , will have
an effect on a very large range of periods, T. An example of this situation is shown in Fig. 4.6,
where mean displacement demands have been computed for a soft soil site with a predominant
period of 1.5 seconds and an effective peak acceleration of 0.25g. In this case, displacement
demands based on elastic analysis can significantly underestimate inelastic demands for buildings
or structures with fundamental periods as long as 1.3 seconds.

43
ll.:nel as tic

fl. elastic
4 ~--------------------------------------------~
ROCK SITES

- - - - - - - 1-L = 1
................ 1-L = 2
-------!-1=3
-----1-L =4
------1-L = 5

0
0.0

Figure 4.1

1.0

0.5

1.5
2.0
PERIOD (sec)

2.5

3.0

Mean of ratio of inelastic to elastic displacement demands for rock sites.

DISPLACEMENT (in)
16
...

......--

/ ...,.---=

!-L=2-

12

___

ROCK SITES

----!-!= 1

14

----~-L=4

10

----1-L

=6

6
4
2

EPA= 0.4g

0.0

Figure 4.2

0.5

1.0

1.5
2.0
PERIOD (sec)

2.5

3.0

Mean of displacement demands in elastic and inelastic systems for rack


sites assuming an EPA of 0.4g.

44
.6.inelastic
.6.elastic

~--------------------------------------------,

ALLUVIUM SITES

\\
3

------
-------------------

1~.

,\~.

\':\
\\ ,.,

......' ...................
''':::;:. ~.

. . . . . .

-~ =--__,~,

!l = 1
!l =2
!l =3
!l = 4
!.1=5
!.1=6

--=-...
------~-

0
0.0

Figure 4.4

0.5

1.0

1.5
2.0
PERIOD (sec)

2.5

3.0

Mean of ratio of inelastic to elastic displacement demands for alluvium


;:;ites.

DISPLACEMENT (in)

16
ALLWIUM SITES

---!.1=1

14

!l = 2

12

----J.L=4

10

- - - - !l = 6

6
4.
2

EPA= 0.4g

0
0.0

Figure 4.3

0.5

1.0

1.5
2.0
PERIOD (sec)

2.5

3.0

Mean of displacement demands in elastic and inelastic systems for alluvium sites assuming an EPA of 0.4g.

45

~inelastic
Llstastic

\\
\\
\

SOFT SOIL SITES

----1-1=1
---------- t.l = 2
- - - - - - - t.l =3
- - - - - t!=4
----t!=S
- - - - - t!=6

\'

\~\ \\_

'-...\
.. \"' "\\'.
'''---

.... -. '~'":=~

... -:..::: ~

0.5

0.0

1.0

1.5

2.0

2.5

3.0

T /Tg
Figure 4.5 Mean of ratio of inelastic to elastic displacement demands for soft soil
sites.

DISPLACEMENT (in)
20
SOFT SOIL SITES

---J.1=1
............... !J. = 2
----!J-=4
----!J-=6

15

10

Tg = 1.5 sec
PGA= 0.25g

0.0

0.5

1.0

1.5

2.0

2.5

3.0

PERIOD (sec)

Figure 4.6 Mean of displacement demands in elastic and inelastic systems for a soft
soil site with a predominant site period of 1.5 sec and assuming a PGA of

0.25g.

47
5. SUMMARY AND CONCLUSIONS
5.1 SUMMARY
Nonlinear response spectra for 124 earthquake ground motions recorded on various soil
conditions, ranging from rock to very soft soils, were computed and analyzed statistically to
'

provide engineers with improved tools to estimate strength and displacement demands on new
and existing buildings. For each record, responses were computed for 50 different periods
between 0.05 and 3.0 seconds, and for 6 displacement ductility ratios, 1, 2, 3, 4, 5 and 6. The
study was limited to computing the responses of SDOF bilinear systems with post-elastic stiffness
of 3% of the elastic stiffness and with a damping ratio of 5% of critical. Average (mean) and
mean plus one standard deviation inelastic strength demand spectra were computed for rock (38
records), alluvium (62 records) and soft soil sites (24 records). These spectra provide adequate
tools with which to estimate strength demands in a deterministic framework.

Based on the results obtained for the strength demand spectra, a comprehensive statistical study
of the strength reduction factor due to hysteretic energy dissipation, R.u , i.e., displacement due
to displacement ductility ratio p., was conducted. The main purpose of this study was to obtain
reliable data on which to judge the reliability of present code recommended values for the
strength reduction factors R and

, and particularly to improve understanding of the factors

affecting these values. Emphasis was given to studying how the values of R14 are affected by soil
conditions, including the effects of very soft soils.

Displacement seismic demand spectra were obtained using the normalized strength demand
spectra.

Present practice for checking against lateral displacement is based on the assumption that the
inelastic displacement demands for severe earthquake ground motions can be based on estimation
of the elastic demands and multiplication of such elastic demands by an empirical coefficient
which is independent of soil conditions.

The reliability of such procedure was studied by

computation of the ratio of inelastic to elastic displacement demands for each of the 124 ground
motions.

48
5. 2 CONCLUSIONS
From analysis of the results obtained, the following observations can be made.

Spectral shapes for inelastic strength demands (p. > 1) differ significantly from elastic

(p.

= 1) spectral shapes.

The largest dynamic amplification for elastic response (p. = 1) is induced by soft soil sites.
However, these large amplifications are significantly reduced when inelastic behavior
occurs (p. > 1). For ductility ratios larger than 3, inelastic strength demands decrease
monotonically with increasing period.

For soft soil records and periods smaller than the predominant period of the site
(T!f8 < 1), there is little difference between the strength demand for ductility ratios
between 2 and 6. This implies that small changes in the yielding strength of structures
with T < T 8 may produce large changes in the ductility demands.

Spectral amplifications between 0.1 and 0.5 seconds for ground motions recorded on soft
soil are usually much smaller than the 2.5 factor which is used to define the Effective
Peak Acceleration (EPA). For this type of soil condition, the use of PGA is probably
more appropriate than the use of EPA.

Normalization of inelastic strength demands using peak acceleration parameters increases


in dispersion with increasing periods.

However, this dispersion was found to be

independent of the ductility level.

Strength reductions due to dissipation of energy induced by hysteretic behavior, i.e.,


R14 , are by no means constant. These reductions (RJ are strongly affected by the natural

period of vibration, the level of displacement ductility ratio, and the local soil conditions.

49

The dispersion of strength reduction factor R.u was observed to be nearly independent of
the period of vibration, and to increase with increasing ductility ratio.

For soft soil conditions, the values of R,u are characterized by small values for
Ttrg < 1 and by very large reductions for periods close to the Tg . The R,u values are

approximately equal to f.l for Ttrg greater than 2.5. This means that estimation of the
predominant period, Tg , of the site . is of particular importance when designing or
upgrading structures on soft soil sites, where inelastic strength demands are strongly
influenced by the Ttrg ratio.

Values of the reduction factor R,u which are based on the assumption that the maximum
displacement of an inelastic system is the same as that for an elastic system, i.e., R,u

=f.l,

are unconservative for structures with short T. Values of R,u which are based on the
assumption that energy absorption in inelastic systems is equal to energy absorption in
elastic systems, i.e., R,u = [2f.l - 1]112, are also unconservative for structures with short T.

The mean values of the ratio of inelastic deformation to elastic deformation show that for
structures with short T the inelastic displacement demands can be considerable larger than
the elastic demands.

The range of the values of the structural period for which elastic analysis can be used
directly to estimate the inelastic displacement demand is dependent on the ductility ratio
level and the soil conditions.

For soft soil sites and for values of Ttrg very near to 1, inelastic displacements can be
up to 40% smaller than the corresponding elastic displacements. For values of Tffg<0.8,
the inelastic displacement demands can be significantly larger than the elastic demands,
so that for sites with very long Tg , the displacement demands based on elastic analysis
can significantly underestimate inelastic displacement demands of structures having T as
large as 1.5 seconds, or even larger, depending on the value of Tg .

50
From the above observations it is obvious that the studies reported herein clearly indicate the
importance of having a reliable estimation of the predominant period of the site, particularly in
case of soft soils.

51

6. REFERENCES
[1]

Bertero, V.V., et al., "Design Guidelines for Ductility and Drift Limits: Review of the
State-of-the-Practice and of-the-Art on Ductility and Drift-Based Earthquake-Resistant
Design of Buildings," A CUREe-Kajima Report, July 1991.

[2]

Krawinkler, H., et al., "Evaluation of Damage Potential of Recorded Ground Motions,"


A CUREe-Kajima Research Report, June, 1991.

[3]

Bertero, V.V., and Uang, C-M., "Issues and Future Directions in the Use of an Energy
Approach for Seismic-Resistant Design of Structures," Proceedings of the Workshop on

Nonlinear Seismic Analysis of RC Buildings, to be published by Elsevier Science


Publishers, October 1991.

[4]

Bertero, V.V., "Structural Engineering Aspects of Seismic Zonation," Proceedings, Fourth


International Conference on Seismic Zonation, Stanford University, August 1991, Vol. 1,
pp. 261-322, Earthquake Engineering Research Institute, Oakland, California.

[5]

Miranda, E., "Seismic Evaluation and Upgrading of Existing Buildings," Ph.D Thesis,
Civil Engineering Dept. of the University of California at Berkeley, California, May,
1991. (To be published in two separate EERC reports).

[6]

Applied Technology Council, "Tentative Provisions for the Development of Seismic


Regulations for Buildings, 11 Report ATC 3-06, National Bureau for Standards, Publication
510, Washington, D.C., June, 1978.

[7]

Seed, H. B., et alia, "Site Dependent Spectra for Earthquake Resistant Design, 11 Report No

EERC 74-12, Earthquake Engineering Research Center, University of California at


Berkeley, California, November, 1974.

52
[8]

Zagajeski, S.W., and Bertero, V.V., "Computer-Aided Optimum Design of Ductile RC


Moment-Resistant Frames," Proceedings, Workshop on Earthquake-Resistant Reinforced
Concrete Building Construction, University of California at Berkeley, California, July
1977, Vol. III, pp. 1140-1174.

[9]

Ridell, R., and Newmark, N.M., "Statistical Analysis of the Response of Nonlinear
Systems Subjected to Earthquakes," Structural Research Series No. 468, Department of
Civil Engineering, University of Illinois, Urbana-Champaign, August 1979.

You might also like