You are on page 1of 16

Proceeding the 6th Civil Engineering Conference in Asia Region: Embracing the Future through

Sustainability
ISBN 978-602-8605-08-3

WHY SHOULD DRIFT DRIVE DESIGN


FOR EARTHQUAKE RESISTANCE?
Mete A. Szen1
1

School of Civil Engineering, Purdue University, West Lafayette, IN. E-mail: sozen@purdue.edu

ABSTRACT
It was the disastrous Messina Earthquake of 1908 that led the structural engineers in Italy to develop
a procedure for earthquake-resistant design based on lateral forces. Considering the physics of structural
response to earthquakes, this decision did not make sense. A structure cannot develop more lateral force
than that limited by the properties of its components. An earthquake shakes a building. It does not load a
building. A building loads itself during a strong earthquake depending on how stiff and strong it is.
Nevertheless, the procedure based on force seemed to work in general. Besides it conformed to the
thinking related to gravity loading and made it convenient to combine effects related to gravity and
earthquake. Admittedly, an engineering design procedure can be good even if it is wrong.
Because it worked, a whole near-science was built around the concept of lateral force. Today, it is not an
exaggeration to claim that the peak ground acceleration is the focal point of almost all that governs
earthquake-resistant design.
In 1932, in a paper not filling a whole page in the Engneering News Record, Harald Westergaard
(Westergaard, 1932) wrote that it was the ground velocity that was the driving factor for damage. His
brilliant insight could have had the profession question whether force was the only issue for design, but it
did not happen.
Over the period 1967-1990, a series of earthquake simulation tests were carried out at the University of
Illinois, Urbana. Although the tests were targeted at the problem of nonlinear dynamic analysis, the most
useful results that emerged were that drift (lateral displacement) was the critical criterion for earthquake
response of a structure, that strength made little difference for the drift response, and that maximum drift
response could be related to peak ground velocity.
The goal of the talk is to explain the changes in thinking inspired by what was observed in the laboratory
and how developments on drift response are likely to affect preliminary proportioning of structures.

INTRODUCTION
There are two simple design rules to achieve satisfactory earthquake resistance of a building structure.
Both rules are related to geometry.
Rule #1: Elevations of the floors must be at approximately the same level after the earthquake that they
were before the earthquake and not as illustrated in Fig. 1. The object of the rule is to save lives.
Rule #2: Geometry of the building on the vertical plane must not differ from its original geometry by
more than a permissible amount on the order of a drift ratio 2 of 1.5% to 2% and not as illustrated in Fig.2.
The object of the rule is to save the investment.

Ratio of lateral relative displacement in a story to the height of the story.

M.A. Szen

Fig. 1: After the Duzce, Turkey, earthquake of 1999


Rule #1 requires adequate detail, such as competent welding for structural steel or the proper amount of
transverse reinforcement for reinforced concrete. It requires a minimum amount of analysis but
considerable amount of knowledge from experience and experiment. The object is to achieve a fail-safe3
structure. Strictly, the engineer does not even need to know the characteristics of the ground motion. All
the engineer needs to know is the finite possibility of the earthquake and be competent in the technology
required to build a fail-safe structure.
Rule #2 requires knowledge of the ground motion to occur and the response of the structure to that ground
motion. Considering the lack of accuracy involved in estimating the ground motion in most localities,
it requires a level of analysis consistent with the expected accuracy of the results in keeping with the tried
and true engineering adage, If one is going to be wrong anyway, one should be wrong the easy way.
This manuscript will focus on Rule #2

Fig. 2: Chile 1985


As it can be inferred from the frequent use of PGA (Peak Ground Acceleration) for rating the intensity of
strong ground motion in engineering design documents, equivalent lateral force tends to be the dominant
driver in determining proportions of structural elements intended to serve as parts of a structural system
resisting earthquake demand. G. W. Housner (Housner, 2002) attributes the use of lateral force in
earthquake resistant design to M. Panetti of Torino Institute of Technology who was a member of the 14person committee formed by the Royal Italian Government after the disastrous earthquake of 28
December 1908. The committee was asked to develop design algorithms to help reduce the damage in
3

Meaning a structure that fails safely and not meaning a structure that does not fail although the no-fail
option would be quite acceptable.

K-17

M.A. Szen

later events. Panetti recognized the need for dynamic evaluation of the entire structure including its
foundation but concluded that this was beyond the state of the art of his time and suggested design for
equivalent static forces (Committee,1909). (Oliveto, 2004) has written that it was S. Canevazzi who
suggested that the committee select buildings observed to have remained intact after the 1908 event and
determine the maximum lateral static forces that they could have resisted, thus providing an observational
basis for specifying lateral force requirements for design. The main concern of the committee members
appears to have been two-story buildings. Studies by the committee resulted in a report requiring design
lateral forces amounting to 1/8 of the upper-story weights and 1/12 of the first-story weight. Inasmuch as
this method was modified after the Tokyo Earthquake of 1923 and re-modified countless times over the
years by different groups in different countries, the main theme did not change. Except for rare instances,
the dominant driver has remained as the equivalent lateral force despite changes in the amount and
distribution of the forces assumed to act at different levels.

A SIMPLE METAPHOR FOR STRUCTURAL RESPONSE TO STRONG


GROUND MOTION

Fig. 3A: Simple metaphor


The response of the simple structural system in Fig. 3 does capture that of a large class of building
structures. The two-dimensional structure in Fig.3a comprises one column fixed to Support A and, a
rigid and strong girder resting on a frictionless roller on Support A. The girder is attached to a large mass.
The flexural response of the column is assumed to be elasto-plastic (Fig. 3b).
If Support A is moved rapidly to the right and if the mass is large enough, the column will develop its
yield moment, My , at both ends (Fig. 3c). It will be subjected to a lateral shear force, V .

(1)
My : limiting moment capacity of column
h : clear column height
V: shear force acting on the column.
We ask a simple question. In the event of the idealized structure being subjected to a sudden horizontal
movement of the foundation, what determines the maximum lateral force on the column?

K-18

M.A. Szen

It is not unreasonable to assume that before the mass will have time to move, the columns will sustain a
lateral deflection that will cause yielding. In that case, the base shear force is going to be that indicated by
Eq. 1. Clearly, the driver is the moment capacity of the column.
Now we need to modify our question. Who determines the base shear?
The answer is the engineer. And that leads us to the next question.
If the base shear is determined by the engineer, except in massive stiff structures, and not by the
earthquake, why do we start the analysis with a crude estimate of the peak ground acceleration? We
should be concerned with the drift and then, if needed, with the lateral force. This hypothetical conclusion
may evoke an objection based on the fact that static design requires the force to determine the drift.
What if the drift can be determined independently of the force or strength in most cases? Is it possible to
determine the drift first and the strength later? Consideration of that possibility is the object of this
discussion.

A BRIEF PERSPECTIVE OF THINKING ON EARTHQUAKE RESISTANT


DESIGN IN THE SEVENTIES
In the 1970s professional thinking on earthquake-resistant design was dominated by at least four strong
currents: (1) the concept of equivalent lateral forces as mentioned above, (2) theoretical considerations
based primarily on experience and interests of mechanical engineers, (3) a response spectrum developed
at the California Institute of Technology translated for design applications as The Los Angeles Formula
that allowed lower base-shear coefficients for taller buildings, and (4) Newmarks stroke of genius in
expressing spectral acceleration response to strong ground motion in three period ranges: (a) nearlyconstant acceleration response, (b) nearly-constant velocity response, and (c) nearly-constant
displacement response4 and then suggesting that energy absorption5 would reduce the calculated linear
response either by the ductility ratio (ratio of yield to displacement capacity of the structure) or by a ratio
based on the energy absorption capacity of the structure, with the reducing factor not exceeding five.
In 1966, a simple experimental system (Sozen and Otani, 1970) to simulate one horizontal component of
earthquake motion was assembled in structural laboratory of the University of Illinois, Urbana (Fig. 4).
The experimental work was initiated by three engineers, Toshikazu Takeda, Polat Glkan, and Shunsuke
Otani Their common goal was a reasonable explanation of the accepted reduction in the force response to
earthquake demand. The first step appeared to be investigation and analytical reproduction of the forcedisplacement history of a reinforced concrete structure in an earthquake. Because almost all of the
thinking about this problem had been in terms of a two-dimensional structure subjected to one horizontal
component of the ground motion, experiments on the Simulator could answer some of the questions.
Takeda had been dealing with the hysteresis problem for reinforced concrete structures at Tokyo
University working with Professor Umemura (Takeda, 1962). At Urbana, he had the opportunity to test
and polish his concept of nonlinear response of reinforced concrete in a dynamic environment.

Much as it has been criticized for not being exact for every kind of ground motion, it is still an effective
way of thinking of the ground-motion demand for proportioning structures.
5

It took some time to change the thinking from blast-resistant design where energy absorption was the
issue to earthquake-resistant design where energy dissipation was the issue.

K-19

M.A. Szen

Fig. 4: Elevation of the university of illinois earthquake simulator


TAKEDA
Using the University of Illinois Earthquake Simulator in 1967, a modest testing machine with a gravityload capacity not exceeding 45 kN and a double-amplitude displacement limit of 0.1 m., Takeda made a
series of simple tests using single-degree-of-freedom specimens to determine the nonlinear response of
reinforced concrete (Fig. 5). The specific object was to determine and formulate the hysteretic response of
reinforced concrete in flexure as it was shaken to develop a series of displacements simulating those that
might occur in an earthquake. The challenge was to calculate the lateral-force and drift response histories
of a reinforced concrete structure throughout the duration of a base motion simulating one horizontal
component of an earthquake motion. Takeda was successful in developing rules for hysteresis of
reinforced concrete in the nonlinear range of response. The focus was on Newmarks concept that the
design force could be expressed as a fraction of the nonlinear-response force depending on the ductility of
the structure. What Newmark accomplished virtually by intuition, Takeda was able to confirm by
calculation based on properties of the structure and the ground motion.

Fig. 5: Test specimen type used by T. Takeda


There was another aspect of Takedas test results that escaped notice, most likely because of the
preoccupation with the magnitude of lateral force. The equivalent lateral force concept, that appeared to
serve well as a design convention, had led to a contradiction between theory and observation. Analysis of
the linear response of an arbitrarily damped structural model resulted in high lateral forces amounting to

K-20

M.A. Szen

multiples of the structural weight, W. On the other hand, observations initiated in 1909, suggested that a
lateral force of approximately W/12 might be sufficient. This created a wide intellectual chasm that
focused attention on force response. The challenge to bring together linear dynamic response, observed
nonlinear response, and a whole host of traditional safety factors led to overlooking Talbots dictum of
observation without preconception.

Fig. 6: Results of test specimen T2 (Takeda)


A simple example is provided by the results of one of Takedas test specimens that was subjected to two
different levels of base motions with the same characteristics in successive test runs. Measured maximum
force and drift responses are summarized in Fig.6 a and b in relation to maximum base acceleration
measured in each test run.
The maximum base acceleration was increased from 1.3g in test run 1 to nearly 2g in test run 2. The
corresponding change in response acceleration was negligible (Fig. 6a). The observed maximum
acceleration response of the mass (directly related to lateral force) was consistent with the concept of
nonlinear (elasto-plastic) concept. Yielding of the column had already occurred in the first test run. Even
though the maximum base acceleration in the second test run was some 60% higher than that in the first
test run, the base shear would not be expected to increase. The maximum response acceleration would
remain the same.
Figure 6b shows the variation of the maximum drift (expressed as a ratio in relation to the column height)
with increase in the maximum base acceleration. The increase in drift ratio of approximately 60% was
comparable to the increase in the base acceleration of approximately 50%. It is, at best, awkward to
explain this result in terms of force begetting drift.
Could it be that drift begets force? This question which could have shortened the time toward a simple
design concept was not asked.

K-21

M.A. Szen

OTANI

Fig. 7: Variation of maximum roof drift with peak base velocity


Otanis main objective was to investigate whether Takedas hysteresis rules could be implemented in
software to determine the response of multi-story reinforced concrete frames to strong ground motion
(Otani, 1972). He developed the required software and tested its results using those from earthquakesimulation of tests of three-story frame structures. While he was doing that he also noted that there was a
linear relationship between maximum drift and maximum base velocity. His penetrating observation
ought to have clinched the idea that there was indeed a direct relationship between base velocity and drift
response, but once again the emphasis on base shear as the driving factor resulted in not appreciating the
importance of his observation.

GLKAN

Fig. 8: Gulkan's definition of substitute damping.


In interpreting the results of single-bay single-story reinforced concrete frames tested using the simulator,
Glkan maximum response in the range of nonlinear response, be it shear or drift, could be explained
with a linear model using his definition of substitute damping in period ranges that would fall within
Newmarks range of nearly constant velocity response and nearly constant displacement response
(Glkan, 1971). His results for substitute damping factors for the tests with base motions simulating
earthquake motions are summarized in Fig. 8.

K-22

M.A. Szen

Fig. 9: Acceleration response spectra at damping factors of 2 and 20%


Glkan observed the substitute damping to increase from 2% of critical at yield to approximately 15% at
a defined ductility factor exceeding 4. Taking an extreme position, one could also interpret the damping
data to suggest a sudden increase to 10% of critical and remain at that level. Linear acceleration response
spectra are shown in Fig. 9 for the two bounding damping levels of 2 and 10 % of critical. Drift response
spectra for the same conditions are shown in Fig. 10.
The professional interest was on the implications of the substitute-damping idealization on force demand.
If a structural system softened (increase in effective period) and developed increased damping under
earthquake demand, the force demand would decrease. To cite a specific example in reference to Fig.9,
if a system with an initial period of 0.5 sec softened to have an effective period of 1 sec. with effective
damping increasing from 2 to 10% of critical, the response acceleration would decrease from 2g to 0.5g
providing a rationale toward explaining the gap between linear-response analysis and design practice.
Figure 10 shows the implications of the substitute-damping idealization on displacement response.
Considering the system with an initial period of 0.5 sec. it is seen that, if the damping increases from 2 to
10 % of critical, the period could increase to 1 sec (a stiffness reduction to of initial) with no increase in
the response displacement. A drift response calculated on the basis of a lightly damped linear system
could be a good estimate of its nonlinear drift response.

Fig. 10: Displacement response spectra at damping factors of 2 and 10 % of critical


ALGAN
In the late seventies Algan (1982) started his work in search of a useful criterion by which to judge the
safety and serviceability of a reinforced concrete structure in a seismic region. It still seems curious that at

K-23

M.A. Szen

that time drift continued to be considered to be a minor consideration in design for earthquake resistance.
This fact is captured very well in Appendix A.
Algans compilation of data on damage revealed that, as long as brittle failure of the structure was
avoided, the story drift ratio (a measure of the distortion of the building profile) was the best pragmatic
indicator of intolerable damage especially because the structure amounted to a fraction of the cost of the
building. His approach demanded a simple and yet realistic procedure for determining drift. To do that he
went back to Glkans substitute damping, by that time expanded by Shibata (Shibata, 1974) into a fullfledged design method. The main conclusion from Algans work was that drift should control design
rather than force and that the main concern for limiting drift was more often related to nonstructural
elements than to the ductility of the structure. This was proposed explicitly during the seventh world
conference on earthquake engineering in 1980 (Sozen,1980).

SHIMAZAKI
In 1982, K. Shimazaki set out in search of an energy based criterion to determine the possible extent of
response-force reduction (Shimazaki, 1984). While pursuing this goal he noticed that within Newmarks
range of nearly-constant velocity response, he could determine the nonlinear drift of reinforced concrete
structures using linear analysis by assuming an amplified period of 2*T where T is the period based on
uncracked state of the structure and a damping factor of 2% of critical. While this was a disturbing
observation because his approach to drift response was insensitive to the area within the hysteresis loop as
well as to strength it confirmed what Fig. 10 implied. Shimazaki reoriented his studies from acceleration
to drift response to come up with a very simple and useful method for determining maximum drift
response. He concluded that if
TR + SR =>1

(2)

DR =<1

(3)

where
TR

: Ratio of T2 to characteristic period for ground motion. The characteristic period was that
beyond
which the spectral energy demand did not increase.
SR
: Ratio of base shear strength to base shear force for linear response
DR
: Ratio of nonlinear-response displacement to linear response displacement based on T2 and
damping factor of 2%.
If the first statement was satisfied, estimating the response displacement was very easy using Eq. 3 but for
stiff structures the method was handicapped.

LEPAGE
In the years 1976 through 1996, a series of earthquake-simulation tests were conducted to produce data on
maximum drift response by With eyes still fixed on force response, experimental work on the simulator
was continued by Aristizabal (1976), Lybas (1977), Healey (1978), Moehle (1978 and 1980), Cecen
(1979), Abrams (1979), Morrison (1981), Kreger (1983), Wolfgram-French (1984), Wood (1985),
Schultz (1985), Bonacci (1989), Eberhard (1989), Dragovich (1996) to investigate the response of sevento ten-story frames and walls (Fig. 9) as well as their interactions. As a result of these works a large
inventory of data on displacement response as well as on modal shapes became available.

K-24

M.A. Szen

Fig. 11: Comparison of Measured and Calculated Drift Maxima.


A study by LePage (1997) to explore the possibility of eliminating the limitation indicated by Eq.2 ended
with a surprisingly simple answer. Using linear models of the test structures and a displacement response
spectrum based on a viscous damping of 2% of critical ,he was able to show that the maximum story
drifts of all structures tested could be closely determined ( Fig. 11). He projected his approach to apply to
building structures subjected to strong ground motion (PGA approximately 0.5g) using a simple
expression relating nonlinear displacement response to building period.

(4)

Sd : nonlinear displacement in m
T : period based on a linear model in sec used as a dimensionless coefficient.

ZTRK
In a study of the effects of the ground motions measured in Anatolia during the two earthquakes of 1999
(Marmara and Dzce), ztrk noted that for structural systems with low base shear strengths and low
mode-1 periods, the drift response tended to exceed the limit set by LePage if the building was
approximately within ten km of the transmitting faulting. He also noted that the difference increased with
peak ground velocity. To generalize what he had observed, he devised Eq. 3.
(5)
Where
(6)

(7)

K-25

M.A. Szen

Sd
PGV
Cy
G
T

: nonlinear spectral displacement


: peak ground velocity
: base shear strength coefficient
: acceleration of gravity
: period in sec. used as a dimensionless coefficient.

Base Shear Strength Coefficient = 0.1


1.2

Spectral Disp., m

1
0.8
0.6
0.4
0.2
0

0.2

0.4

0.6

0.8

1.2

1.4

1.6

1.8

Period, sec.

Fig. 12A: Variation of nonlinear displacement response with period for a base shear
strength coefficient of 0.1
The results of Eq. 4 are shown in Fig. 12a and 12b for base shear strength coefficients of 0.1 and 0.2 and
for peak ground velocities of 0.4 m/sec (broken line) and 0.8 m/sec (solid line). It is seen that as the base
shear strength coefficient increases, the maximum response spectrum tends to conform to the LePage
spectrum over a larger range of periods. In effect, ztrk had returned in 2003 to what Westergaard had
suggested in 1932.

Base Shear Strength Coefficient = 0.2

Spectral Disp., m

1.2
1.1
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0

0.2

0.4

0.6

0.8

1.2

1.4

1.6

1.8

Period, sec.

Fig. 12b: Variation of nonlinear displacement response for a base shear strength
coefficient of 0.2

K-26

M.A. Szen

CONCLUSION
It is important to emphasize that the drift determination is useful in comparing and selecting among
competing design options. One should not expect to obtain an exact drift prediction, not only from the
procedure described here, but from any procedure that is proposed for calculation of drift caused by an
earthquake that has not happened. Within that understood, it is reasonable to choose the right structural
framing for building structures of moderate height (below 20 stories) using the following approach.
For the purpose of choosing between competing designs, the spectral drift may be estimated from the
expression

(4)

where
Sd
T
code)

: Spectral displacement in m
: Computed mode-1 period of structure (dimensionless and not a period based on tradition or

: A constant that may be taken as 4 for a strong ground motion expected to have a peak ground
acceleration of 0.5g +/- 0.1g and a velocity response that may be assumed to remain nearly
constant over the period range 0.5 to 2.5 sec.
If the site of the structure is suspected to be within 10 km of a transmitting fault and if its first-mode
period does not exceed 1 sec, the spectral drift may be estimated using Eq. 5
If the calculated first-mode period of the structure in question does not exceed 2.5 sec, story drifts in a
multi-story structure may be estimated using the expression

mi i

i 1

m i i
n

i S d

(5)

i 1
Di
n
mi
i

: Drift at level i
: Number of stories
: Mass at level i
: modal shape factor at level i (it is convenient to define the modal shape by normalizing
displacements at all levels with respect to the displacement at top level)
Sd
: Spectral displacement
Equation 5 is useful if the distribution of mass and stiffness over the height of the structure is not uniform.
Otherwise, selection of the framing with the correct proportions may be made using Eq. 4.

NOTE
All Structural Research Series reports cited below may be downloaded from
http://www.ideals.illinois.edu/
Complete sets of data from many of the experiments reported in the reports listed below may be found in
https://nees.org/warehouse/enhanced

K-27

M.A. Szen

REFERENCES
Abrams, D.P. (1979). Experimental Study of Frame-Wall Interaction in Reinforced Concrete Structures
Subjected to Strong Earthquake Motions. Civil Engineering Studies, Structural Research Series No. 460,
University of Illinois, Urbana, Il., May 1979, 386 pp.
Algan, B. (1982). Drift and Damage Considerations in Earthquake-Resistant Design of Reinforced
Concrete Buildings, PhD Thesis Submitted to The Graduate College of The University of Illinois,
Urbana, IL, 1982.
Aristizabal-Ochoa, Dario (1976). Behavior of Ten-Story Reinforced concrete Walls Subjected to
Earthquake Motion. Civil Engineering Studies, Structural Research Series No. 3431, University of
Illinois, Urbana, Il., October 1976, 5378 pp.
Committee Report, (1909). Norme Edilizie Obligatorie per I Comuni Colpiti dal Terremoto del 28
Diciembre 1908 e da altri anteriori, (Mandatory Building Codes for Municipalities Affected by The
Earthquake of 28 December 1908 and Previous Events) Giornale del Genio Civile, Roma, 1909.
Committee on Special Structures, Structural Engineers Association of Southern California, Report on
Drift, Los Angeles, CA, 1959.
een, Haluk, (1979). Response of Ten-Story Reinforced Concrete Model Frames to simulated
Earthquakes. PhD Thesis Submitted to the Graduate College of The University of Illinois, Urbana, IL,
June 1979.
Dragovich, J.J. (1996). An Experimental Study of Torsional Response of Reinforced Concrete Structures
to Earthquake Excitation. PhD Thesis Submitted to the Graduate College of The University of Illinois,
Urbana, IL, June 1996.
Glkan, Polat, (1971). Response and Energy-Dissipation of Reinforced Concrete Frames Subjected to
Strong Base motions. Civil Engineering Studies, Structural Research Series No. 377, University of
Illinois, Urbana, Il., May 1971, 288 pp.
Healey, Timothy, (1978). Experimental Study of The Dynamic Response of A Ten-Story Reinforced
Concrete Frame with A Tall First Story. Civil Engineering Studies, Structural Research Series No. 450,
University of Illinois, Urbana, Il., August 1978, 120 pp.
Housner, G.W. (2002). Historical View of Earthquake Engineering. International Handbook of
Earthquake & Engineering Seismology, Part A, Volume 81A , 2002, pp 13-18.
Kreger, M. (1983). A Study of Causes of Column Failures in The Imperial County Services Building
during The 15 October 1979 Imperial Valley Earthquake. Civil Engineering Studies, Structural Research
Series No. 509, University of Illinois, Urbana, Il., August 1983, 321 pp.
Lepage, Andres, (1996). A Method for Drift Control in Earthquake-Resistant Design of Reinforced
Concrete Building Structures. PhD Thesis Submitted to the Graduate College of The University of
Illinois, Urbana, IL, October 1996.
Lybas, John (1977). Effect of Beam Strength and Stiffness on Dynamic Behavior of Reinforced Concrete
Coupled Walls. Civil Engineering Studies, Structural Research Series No. 392, University of Illinois,
Urbana, Il., July 1977, 569 pp.
Moehle, J. P. (1978). Earthquake Simulation Tests of A Ten-Story Reinforced Concrete Frame wit A
Discontinued First-Level Beam. Civil Engineering Studies, Structural Research Series No. 451,
University of Illinois, Urbana, Il., August 1978, 162 pp.
Moehle, J. P. (1980). Experiments to Study Earthquake Response of Reinforced Concrete Structures with
Stiffness Interruptions. Civil Engineering Studies, Structural Research Series No. 482, University of
Illinois, Urbana, Il., August 1980, 542 pp.

K-28

M.A. Szen

Morrison, D.G. (1980). Response of Reinforced Concrete Plate-Column Connections to Dynamic and
Static Horizontal Loads. Civil Engineering Studies, Structural Research Series No. 482, University of
Illinois, Urbana, Il., August 1980, 217 pp.
Oliveto, G. (2004). Review of the Italian Seismic Code Released after The 1908 Messina Earthquake.
Proceedings of the Passive Structural Control Symposium, Tokyo Institute of Technology, November
2004, pp. 1-20.
Otani, Shunsuke (1972). Behavior of Multistory Reinforced Concrete Frames During Earthquakes. Civil
Engineering Studies, Structural Research Series No. 392, University of Illinois, Urbana, Il., November
1972, 575 pp.
ztrk, Baki, (2003). Seismic Drift Response of Building Structures in Seismically Active and Near-Fault
Regions. PhD Thesis submitted to the Faculty of Purdue University, May 2003.
Saiidi, Mehdi (1979). Simple and Complex Models for Nonlinear Seismic Response of Reinforced
Concrete Structures. Civil Engineering Studies, Structural Research Series No. 465, University of Ill.,
Urbana, Il., August 1979, 188 pp.
Schultz, A.E. (1985). An Experimental and Analytical Study of Reinforced Concrete Frames with Yielding
Columns. PhD Thesis Submitted to the Graduate College of The University of Illinois, Urbana, IL, Augus
1985.
Shibata, Akenori(1974). The Substitute-Structure Method for Earthquake resistant Design of Reinforced
Concrete Frames. Civil Engineering Studies, Structural Research Series No. 412, University of Illinois,
Urbana, Il., October 1974, 34 pp.
Shimazaki, K. (1984). Seismic Drift of Reinforced Concrete Structures. Technical Report, Hazama-Gumi,
Tokyo, 1984. P. 145-165.
Szen, M.A. (1980). Review of Earthquake Response of Reinforced Concrete Buildings with A View to
Drift Control. Seventh World Conference of Earthquake Engineering, Istanbul, Turkey, 1980.
Szen, M.A. and Otani, S. (1970). Performance of The University of Illinois Earthquake Simulator for
Reproducing Scaled Earthquake Motions. Proceedings of US-Japan Seminar in Erathquake Engineering,
Sendai, Japan, September, 1970.
Takeda, Toshikazu (1962). Study of the load-Deflection Characteristics of Reinforced Concrete Beams
Subjected to Alternating Loads. Transactions, Architectural Institute of Japan, V. 76, 1962.
Takeda, Toshikazu et al. (1970). Reinforced Concrete Response to Simulated Earthquakes. Journal of the
Structural Division, American Society of Civil Engineers, No. ST12, December, 1970, p. 2557- 2573.
Eberhard, Marc (1989). Experiments and Analyses to Study The Seismic Response of Reinforced Concrete
Frame-Wall Structures with Yielding Columns. Civil Engineering Studies, Structural Research Series No.
548, University of Illinois, Urbana, Il., September 1989, 5424 pp.
Bonacci, J..F. (1989). Experiments to Study Seismic Drift of Reinforced Concrete Structures. PhD Thesis
Submitted to the Graduate College of The University of Illinois, Urbana, IL, June 1989.
Severn, R.T. (2010). The Development of Shaking TablesA Historical Note. Earthquake Engineering
and Structural Dynamics, 40, 2010, p 195-213.
Veterans Administration Office of Construction, Report of the Earthquake and Wind Forces Committee
(1972). Earthquake-Resistant Design Requirements for VA Hospital Facilities.s Washington, D.C.1972.
Westergaard, H.M. (1932). Measuring Earthquake Intensity in Pounds per Square Foot. Engineering
News Record.

K-29

M.A. Szen

Wolfgram-French, Catherine, E. (1984). Experimental Modeling and Analysis of Three One-Tenth Scale
Reinforced Concrete Wall-Frame Structures, PhD Thesis Submitted to the Graduate College of The
University of Illinois, Urbana, IL, June 1984, 350 pp.
Wood, Sharon, L. (1985). Experiments to Study the Earthquake Response of Reinforced Concrete Frames
with Setbacks. Civil Engineering Studies, Structural Research Series No. 544, University of Illinois,
Urbana, Il., December 1985, 390 pp.

K-30

M.A. Szen

APPPENDIX A
REPORT ON DRIFT
COMMITTEE ON SPECIAL STRUCTURES
STRUCTURAL ENGINEERS ASSOCIATION OF SOUTHERN
CALIFORNIA
This committee met with the staff of the Building Division of the Department of Building and Safety on
Friday evening, February 27, 1959. Present were Murray Erick, Robert Wilder, Roy Johnston, and S. B.
Barnes. R. W. Binder was absent but sent in his written comments. These were read at the meeting and
given serious consideration.
The meeting was held in the Hearing Room at the City Hall after dinner at the Redwood House.
The question submitted to the committee was: "What is good engineering practice as related to drift of
buildings subjected to lateral forces?" The first part of the meeting was devoted to a discussion of the
extent to which a building department should set limitations of this kind and to the desirability of
attempting to legally limit the amount of nonstructural damage. The necessity of using "sound
engineering judgment", that overworked phrase, was stressed. Three items were discussed under the
category of basic philosophy, namely-protection of health and safety of life, protection of nonstructural
elements, and protection from motion sickness or discomfort.
It was felt by the committee that the breakage of glass or the breaking up of exterior wall elements which
might fall on the public was definitely within the province of regulation. The committee generally felt that
the other two items bordered on excessive paternalism. Mr. Hanley Wayne disputed this point of view on
the basis that the present law required them to safeguard property. It was pointed out that some plaster
cracks must be expected in case of high wind or severe earthquake even with the most conservative
designs. It was noted that isolation of plaster walls, glass or other brittle materials was possible in some
degree but that more isolation would result in less damping. It was finally decided that we could not
legislate out all plastering cracks but that it would be desirable to avoid breaking up partitions beyond the
point of mere plaster repair. The chairman started the meeting by questioning whether this matter
rightfully should come before this committee. He personally feels that this should be a matter for the
entire Seismology Committee since it involves general principles. However, the committee proceeded
with its discussion and findings, hoping that it was not too far out of line in this respect.
It was further recommended that the advice and opinions of the committee on this matter not be placed in
code form but be considered as a portion of a Manual of Good Practice, which could be revised later if
found desirable. Since most experience of record as related to drift in tall buildings has been due to the
effect of wind and since such recommendations of experienced engineers are limited to wind, it was
decided to make separate recommendations for wind and for earthquake. It was felt by most of the
committee that the deflections of buildings subjected to earthquake computed as a statical force are not
necessarily a true measure of the actual deflections of these buildings under earthquake shock. It was
therefore decided to permit earthquake drift to be twice that for wind. The monetary cost of extreme
limitation for earthquake drift was discussed and considered in this recommendation. Because it is
desirable to have uniform agreement between engineers on a state-wide basis, and because the SEAOC is
currently proposing a new set of earthquake design criteria to the Uniform Code officials, and since there
is at present a committee writing a Manual of Good Practice which will involve drift limitations, it is
suggested that this report be sent to the proper committees engaged in this work for their review and
comments.
An excerpt of conclusions reached at the meeting, compiled by Tom Brown and Hanley Wayne, is
enclosed.
Respectfully submitted,
COMMITTEE ON SPECIAL STRUCTURES,
SEAOSC
S. B. Barnes, Chairman
Murray Erick
Roy Johnston
Robert Wilder
R. W. Binder
Los Angeles, California
March 4, 1959

K-31

You might also like