Sie sind auf Seite 1von 16

Proceedings of Fourth Turkish National Conference on Earthquake

Engineering, Ankara, Turkey, September 17-19, 1997

LESSONS LEARNED FROM PAST EARTHQUAKES

Shunsuke Otani

ABSTRACT

The development of design seismic forces is reviewed. The earthquake engineering was
shown to have developed in the last 60 years. The characteristics of ground motion and
structural response were gradually understood over the last five decades. The damage
statistics from recent major earthquakes indicated that the severe damage was observed
in taller buildings. The ratios of buildings with operational damage were as high as 90
percent because predominant buildings were low-rise buildings and such low-rise
buildings suffered much less damage. Contrary to the statistics, the nonlinear response
analysis results showed higher ductility demand (damage) to low-rise buildings. Causes
of earthquake damage are briefly discussed for reinforced concrete buildings. A large
number of problems have been solved in research community. It is essential for the
disaster reduction that the information in research domain should be transferred to the
engineering community and implemented in engineering practice. The future direction
of earthquake engineering is discussed in terms of structural engineering.

INTRODUCTION

Engineering appears similar to Science, but there is an obvious difference between


the two. The objective of engineering is to produce a product to meet people's demand,
whereas the objective of science is to understand the natural phenomena. The theoretical
soundness and basis are essential in science. However, they may be desired in
engineering, but are not essential. A large error could be tolerated if sound engineering
judgement can be made.
Earthquake engineering is aimed to build human environment safe from an
earthquake disaster. Therefore, our predecessors studied each case of earthquake
disasters carefully immediately after the event, and investigated possible causes and
their solutions. Counter measures have been developed against thousands of bitter
experiences to mankind from earthquake disasters. As this engineering has been built on
the experience, it is not theoretically consistent. Although the earthquake engineering
has improved the built environment in the world, there remains a lot to be improved to
protect human life from future earthquake disasters.
This paper concentrates on lessons learned from past earthquakes about earthquake
resistance of buildings with an emphasis on reinforced concrete structures.

___________________________
Professor, University of Tokyo, Department of Architecture, Tokyo, Japan

1
EQUIVALENT STATIC SEISMIC DESIGN FORCES

Dr. G. W. Housner presented a conference lecture, entitled "Historical View of


Earthquake Engineering [1]," at the opening of Eighth World Conference on Earthquake
Engineering (1984) in San Francisco. He pointed out that "Earthquake engineering is a
20th Century development, so recent that it is yet premature to attempt to write its
history." As an example, the development of seismic design forces is reviewed below.
J. Milne (1850-1913), an invited British professor in mining and geology at the
College of Engineering, Imperial University of Japan (University of Tokyo), recorded
the disaster of the 1891 Nohbi (Mino-Owari) Earthquake in Japan [2]. The earthquake
(M 7.4) killed more than 7,000 people and caused a significant damage to then modern
(western) brick construction in Nagoya area. Milne [2] reported that "buildings on soft
ground … suffer more than those on the hard ground." and that "… we must construct,
not simply to resist vertically applied stresses, but carefully consider effects due to
movements applied more or less in horizontal directions." The resistance against
horizontal forces was already recognized as necessary for structures, but the
quantification of lateral forces had not been formulated in the last century. After this
earthquake the study for earthquake resistant construction started in Japan, and the
improvement of structural detailing was recommended for timber construction.
Quantitative definition of design seismic forces was proposed after the 1908
Messina Earthquake (M 7.5), Italy, which killed more than 83,000 people. Dr. Housner
states [1] that "The government of Italy responded to the Messina earthquake by
appointing a special committee composed of nine practicing engineers and five
professors of engineering to study the earthquake and to make recommendations. … M.
Panetti, Professor of Applied Mechanics in Turin ... recommended that the first story be
designed for a horizontal force equal to 1/12 the weight above and the second and third
stories to be designed for 1/8 of the building weight above." Although this was the first
attempt to quantify the design earthquake force, this recommendation was made without
the knowledge about the earthquake ground motion nor the understanding of the
structural resistance.
The 1923 Kanto (Tokyo) Earthquake (M 7.9) caused significant damage in Tokyo
and Yokohama areas, Japan, with a loss of life more than 140,000. More than 90 percent
of the dead and lost were caused by fire. The Urban Building Law Enforcement
Regulations introduced a seismic coefficient of 0.10 in the seismic design requirements
in 1924; i.e., maximum ground acceleration of 0.3 g estimated in downtown Tokyo
during the Kanto Earthquake was divided by the safety factor of 3 used in determining
allowable stress level. No response amplification by a structure was considered. In the
U.S., the 1925 Santa Barbara Earthquake, California, triggered the adoption of
earthquake resistant building code in several Californian communities. The first edition
of the Uniform Building Code in 1927 suggested a lateral base shear coefficient of 0.1
for buildings in high seismic risk areas.
Although design seismic forces were specified in the code, no practical methods of
stress analysis were made available to engineers. A structural analysis method available
at the time was, for example, Castigliano's theorem, but such were tools only for
academic researchers. An inaccurate portal method was used by engineers when Dr. H.
Cross [3] introduced the moment-distribution method in 1930. Dr. K. Muto [4] proposed
the D-value method for a frame analysis (a practical method for design using tables and
figures) in 1933.

2
Indeed, it is not a long time ago that our predecessors were provided with
minimum tools to calculate the stresses in a structure under a simulated earthquake
effect.

EARTHQUAKE MOTION AND STRUCTURAL RESPONSE

Dr. Housner pointed out [1] that "The first accelerographs were installed by the
Seismological Field Survey of the U.S. Coast and Geodetic Survey in late 1932, just in
time to record the strong ground shaking of the destructive March 10, 1933 Long Beach
earthquake. This was a most important step in the development of earthquake
engineering." Since then, many accelerographs have been installed in the ground and on
the structure. These records have provided us with precious information about the
characteristics of earthquake motions and structural responses. A strong ground motion,
recorded at El Centro station during the 1940 Imperial Valley Earthquake, has been
treated as a standard strong motion and has been used in research as well as in design
practice.
The impact of theoretical analysis was felt for the first time in earthquake
engineering when the earthquake response was calculated by a mechanical response
analyzer [5]. The fact that more flexible buildings would typically be subjected to lower
seismic forces and that the lateral force distribution would not be uniform over the
structural height was recognized in the 1943 building code provisions of the City of Los
Angeles. The seismic probability map was published by the U.S. Coast and Geodetic
Survey in 1948 and was adopted in the 1952 Uniform Building Code. In this manner,
the concept of probability was introduced in determining seismic risks.
The development of analog as well as digital computers accelerated the
understanding of structural response under earthquake excitations. Nonlinear earthquake
response studies of simple systems by A. S. Veletsos and N. M. Newmark [6] provided
guidelines to estimate lateral force resistance required for a system having a specified
deformation capability; i.e., (a) equal energy rule for short- to medium-period systems,
and (b) equal displacement rule for long-period systems. This concept was used to
rationalize the design seismic forces being much smaller than the elastic response
spectral acceleration. The 1959 SEAOC Code [7] introduced the type of construction
factor to recognize the ductility of a structural system and the importance factor for the
use of a building.
The accumulation of experimental data on the behavior of structural models and
members under simulated earthquake forces made it possible to analyze a realistic
inelastic response of structures during an earthquake in the late 1960s.
J. Milne observed the difference in building response on firm and soft soils at the
end of the last century [2]. The effect of surface geology on ground motion
amplification was recognized in the 1976 ATC-03 recommendation [8].

EARTHQUAKE DAMAGE STATISTICS

Reliable statistics about damaged and undamaged buildings are important to


establish disaster reduction measures against future earthquakes. Some comprehensive

3
statistics were obtained after the 1985 Mexico Earthquake, the 1990 Luzon
(Philippines) Earthquake, the 1992 Erzincan (Turkey) Earthquake and the 1995 Hyogo-
ken Nanbu (Kobe) Earthquake.
The damage of all buildings in a selected area was surveyed by external
observation. The damage was grouped into three levels (six levels in the original
investigations); i.e., (a) operational damage: columns or structural walls were slightly
damaged in bending, and some shear cracks might be observed in non-structural walls,
(b) heavy damage: spalling and crushing of concrete, buckling of reinforcement, or
shear failure in columns were observed, and lateral resistance of shear walls might be
reduced by heavy shear cracking, and (c) collapse, which also included those buildings
demolished at the time of investigation.

The 1985 Mexico Earthquake

An M8.1 earthquake occurred on the Mexican west coast on September 19, 1985,
followed by an M7.5 after-shock on September 21. The two successive events caused a
significant damage in Mexico City approximately 400 km away from the epicenter. The
structural damage was concentrated in a region of old lake bed which was gradually
filled with the development of the city. The damage was attributed to the amplification
of long-period components of the ground motion by deep and soft soil deposit underlain
in the Mexico Valley.
The damage statistics in Table 1 were obtained one to two months after the
earthquake in the old lake bed zone, including heavily damaged residential as well as
commercial districts, covering slightly more than 20 percent of the central city [9].
Approximately 73.4 percent of the 4,532 buildings surveyed were three-stories or lower;
94.0 percent of buildings were six-story or lower.
Ninety-four (93.8) percent of the buildings were judged operational in the
surveyed area because more than 94 percent of those dominantly low-rise buildings (six
stories or lower) were operational. The damage increased sharply in mid- to high-rise
buildings (seven stories and higher). Nearly or more than half of nine-story and taller
buildings suffered heavy damage or collapsed; 37.2 percent out of 129 buildings
suffered heavy damage and 15.5 percent collapsed. The ground motion, dominated by
long period components in the lake bed zone, definitely increased the damage of taller
buildings. The 19-story building with operational damage is the Latin Americana Tower
located at the center of a commercial district.
The damage was light in a area where stiff low-rise buildings were dominant. It
should be noted that the percentage of severe damage was extremely small up to six-
story buildings in comparison with those in the other statistics. The heavy damage and
collapse of tall buildings gave a significant impact to the society. The flat slab
construction collapsed in a form of piled pancakes, leaving no space for life survival
between adjacent slabs. The pounding of adjacent buildings was observed partially due
to the rocking motion of the structures.
Tables 2 and 3 show the damage of reinforced concrete and masonry buildings in
the city of Lazaro Cardenas, in the epicenter region. The damage in one- and two-story
reinforced concrete buildings was relatively small, but severe damage increased with
the number of stories, especially in four-story or taller buildings. The damage to one-
and two-story masonry construction was also light.

4
Table 1: Damage statistics of buildings in lake bed zone of Mexico City [9]
No. of Operational Heavy
Stories Damage Damage Collapse Total
1 689(97.5) 17(2.4) 1(0.1) 707(100)
2 1,756(96.0) 57(3.1) 17(0.9) 1,830(100)
3 752(95.3) 26(3.3) 11(1.4) 789(100)
4 412(95.2) 16(3.7) 5(1.2) 433(100)
5 332(94.1) 9(2.5) 12(3.4) 353(100)
6 140(94.0) 2(1.3) 7(4.7) 149(100)
7 74(75.5) 11(11.2) 13(13.3) 98(100)
8 34(77.3) 9(20.5) 1(2.3) 44(100)
9 23(54.8) 8(19.0) 11(26.2) 42(100)
10 15(62.5) 6(25.0) 3(12.5) 24(100)
11 11(57.9) 8(42.1) 0(0.0) 19(100)
12 3(18.8) 10(62.5) 3(18.8) 16(100)
13 1(14.2) 6(85.7) 0(0.0) 7(100)
14 2(50.0) 2(50.0) 0(0.0) 4(100)
15 4(44.4) 5(55.5) 0(0.0) 9(100)
16 0(0.0) 1(50.0) 1(50.0) 2(100)
17 2(40.0) 1(20.0) 2(40.0) 5(100)
18 1(100) 0(0.0) 0(0.0) 1(100)
Total 4,251(93.8) 194(4.3) 87(1.9) 4,532(100)
( ) : percentage to the total of the same story height.

Table 2: Damage statistics of RC buildings in Lazaro Cardenas [9]


No. of Operational Heavy
Stories Damage Damage Collapse Total
1 45(90.0) 5(10.0) 0(0.0) 50(100)
2 71(89.9) 7(8.9) 1(1.3) 79(100)
3 13(72.2) 4(22.2) 1(5.6) 18(100)
4 6(50.0) 6(50.0) 0(0.0) 12(100)
5 2(50.0) 2(50.0) 0(0.0) 4(100)
9 0(0.0) 1(100) 0(0.0) 1(100)
Total 137(83.5) 25(15.2) 2(1.2) 164(100)
( ) : percentage to the total of the same story height.

Table 3: Damage statistics of masonry buildings in Lazaro Cardenas [9]


No. of Operational Heavy
Stories Damage Damage Collapse Total
1 82(100) 0(0.0) 0(0.0) 82(100)
2 24(88.9) 2(7.4) 1(3.7) 27(100)
3 0(0.0) 1(100) 0(0.0) 1(100)
Total 106(96.4) 3(2.7) 1(0.9) 110(100)
( ) : percentage to the total of the same story height.

5
The 1990 Luzon (Philippines) Earthquake

An earthquake (M7.7) occurred approximately 100 km to the north of Manila,


Philippines, on July 16, 1990, and killed approximately 2,000 people. Many high-rise
hotel buildings collapsed in the resort city of Baguio. The damage statistics were
collected in a major commercial district of the city of Baguio. Ninety-three (92.8)
percent of the buildings in the area were of five stories or lower.
Seventy-six (76.2) percent of the buildings suffered operational damage, 18.8
percent suffered heavy damage, and 5.0 percent collapsed. Severe damage was observed
in taller buildings. The percentage of heavy damage and collapsed buildings was higher
in the low-rise buildings (six-story or lower) compared with the other earthquake cases.

Table 4: Damage statistics of RC buildings in Baguio City [10]


No. of Operational Heavy
Stories Damage Damage Collapse Total
1 5(71.4) 2(28.6) 0(0.0) 7(100)
2 34(75.6) 10(22.2) 1(2.2) 45(100)
3 42(84.0) 7(14.0) 1(2.0) 50(100)
4 36(85.7) 4(9.5) 2(4.8) 42(100)
5 15(62.5) 7(29.2) 2(6.9) 24(100)
6 4(80.0) 1(20.0) 0(0.0) 5(100)
7 0(0.0) 2(50.0) 2(50.0) 4(100)
8 2(66.7) 0(0.0) 1(33.3) 3(100)
9 0(0.0) 1(100) 0(0.0) 1(100)
Total 138(76.2) 34(18.8) 9(5.0) 181(100)
( ) : percentage to the total of the same story height.

The 1992 Erzincan Earthquake

An Ms6.9 earthquake occurred on March 13, 1992, with an epicenter (focal depth
of 28 km) near the City of Erzincan, in the eastern part of the Anatolia Plateau, Turkey.
In the City of Erzincan, two heavily damaged residential areas were selected for the
inventory survey or reinforced concrete buildings; i.e., (a) Fatih, Yunus and Aksemsettin
Districts, and (b) Yavus Selim District. Ninety (90.3) percent of the buildings in the area
were 2- to 4-story apartment buildings. A joint team, organized by the Architectural
Institute of Japan and the Japan Society for Civil Engineers, worked with the
researchers from Bogazici University, Istanbul [11]. Dr. P. Gulkan, Middle East
Technical University, provided guiding information about the damage in the area.
The damage of one- and two-story buildings was very light (Table 5). No three-
story buildings collapsed, but 23 percent of the buildings suffered heavy damage.
Seventeen (17.3) percent of four-story buildings collapsed and 30.9 percent suffered
heavy damage.
It is interesting to note that the building height was limited to three stories along
main streets and two stories in the other area after the devastating 1939 Erzincan
Earthquake (M 7.9). A recent growth in population and economy as well as strong
confidence in technical development relaxed the height limitation; i.e., six-story
buildings are allowed along main streets and four-story buildings in the other areas.

6
Damage must have been much smaller without the relaxation in building height.

Table 5: Damage statistics of reinforced concrete buildings in Erzincan [11]


No. of Operational Heavy
Stories Damage Damage Collapse Total
1 28(100) 0(0.0) 0(0.0) 28(100)
2 125(100) 0(0.0) 0(0.0) 125(100)
3 114(77.0) 34(23.0) 0(0.0) 148(100)
4 57(51.8) 34(30.9) 19(17.3) 110(100)
5 4(100) 0(0.0) 0(0.0) 4(100)
unknown 0(0.0) 0(0.0) 9(100) 9(100)
Total 328(77.4) 68(16.0) 28(6.6) 424(100)
( ) : percentage to the total of the same story height.

The 1995 Hyogo-ken Nanbu (Kobe) Earthquake

The 1995 Hyogo-ken Nanbu Earthquake killed nearly 6,300 people (including
direct and indirect causes by the earthquake), injured more than 40 thousands people.
The Kinki (Osaka and Kyoto Region) Branch, Architectural Institute of Japan,
investigated the damage of all existing reinforced concrete buildings (3,911 buildings in
total) in the region of seismic intensity VII (highest seismic intensity defined by the
Japan Meteorological Agency) in Nada and Higashi-Nada wards, Kobe City [12].
Seventy-five (75) percent of the 3,911 buildings surveyed were residential
buildings (including those used partially for office or shop). Forty-eight (47.5) percent
of the buildings were built in conformance with the current Building Standard Law
(revised in 1981). Eighty-three (82.5) percent of the buildings were lower than or equal
to five stories high (73.0 percent were from 3 to 5 stories high) and 9.7 percent had soft
first-story construction.
Eighty-nine (88.5) percent of the buildings surveyed suffered operational damage,
5.9 percent suffered heavy damage and 5.7 percent collapsed. Among those 2,035
buildings constructed before the current Building Standard Law (1981), 7.4 percent
suffered heavy damage and 8.3 percent collapsed. Among those 1,859 buildings
constructed using the current Building Standard Law, 3.9 percent suffered heavy
damage and 2.6 percent collapsed. The 1981 revision of the Building Standard Law
enhanced significantly the performance of reinforced concrete buildings against
earthquake attack.
The relation between damage level and the number of stories is shown in Tables 6
and 7 for buildings constructed before and after the enforcement of the 1981 Building
Standard Law. The ratio of severer damage increases with the number of stories,
especially for buildings taller than 5 stories. The percentage of the pre-1981 buildings,
which remained operational after the earthquake, is 87.9 percent for 1- to 5-story
buildings, 66.1 percent for 6- to 8-story buildings, and 46.0 percent for 9-story and
higher buildings. The corresponding percentages of the post-1981 buildings are 97.1
percent, 87.1 percent and 70.7 percent, respectively. The revision of the law improved
the level of safety almost uniformly from low-rise to mid-rise buildings. Even after the
revision of the law, 20 percent of buildings taller than seven stories suffered heavy
damage or collapsed.

7

Table 6 Damage of RC buildings in Kobe constructed before 1981 [12]
No. Operational Heavy
Stories Damage Damage Collapse Total
1 20(90.9) 1( 4.5) 1( 4.5) 22(100)
2 215(92.7) 9( 3.9) 8( 3.4) 232(100)
3 532(93.0) 17( 3.0) 23( 4.0) 572(100)
4 524(85.8) 41( 6.7) 46( 7.5) 611(100)
5 269(79.6) 29( 8.6) 40(11.8) 338(100)
6 59(75.6) 10(12.8) 9(11.5) 78(100)
7 49(58.3) 16(19.0) 19(22.6) 84(100)
8 19(63.3) 7(23.3) 4(13.3) 30(100)
9 3(33.3) 4(44.4) 2(22.2) 9(100)
10 20(48.8) 15(36.6) 6(14.6) 41(100)
( ): Ratio to the number of buildings of the same height(%)


Table 7 Damage of RC buildings in Kobe constructed after 1981 [12]
No. Operational Heavy
Stories Damage Damage Collapse Total
1 8( 100) 0( 0.0) 0( 0.0) 8(100)
2 85(98.8) 0( 0.0) 1( 1.2) 86(100)
3 460(98.1) 2( 0.4) 7( 1.5) 469(100)
4 508(95.7) 9( 1.7) 14( 2.6) 531(100)
5 333(97.4) 5( 1.5) 4( 1.2) 342(100)
6 135(91.8) 9( 6.1) 3( 2.0) 147(100)
7 90(86.5) 12(11.5) 2( 1.9) 104(100)
8 44(75.9) 11(19.0) 3( 5.2) 58(100)
9 19(73.1) 7(26.9) 0( 0.0) 26(100)
10 51(69.9) 18(24.7) 4( 5.5) 73(100)
( ): Ratio to the number of buildings of the same height(%)

DESIGN SEISMIC FORCES AND EARTHQUAKE RESPONSE

A series of single-degree-of-freedom (SDF) systems were designed using the


governing building code. The nonlinear response of the SDF systems was calculated
under earthquake motions recorded near the area of the inventory damage surveys after
the 1985 Mexico Earthquake and the 1992 Erzincan Earthquake. The Takeda model
[13] was selected to simulate the response of well-designed reinforced concrete
buildings. The skeleton curve was idealized by a tri-linear relationship with stiffness
changes at cracking point (Dc, Fc) and yielding point (Dy, Fy). Fixed relations were
arbitrarily assumed for the cracking and yielding points; Fc = Fy / 3 and Dc = Dy / 12.
Mass of the system was assumed to be unity. The damping coefficient was assumed to
vary proportional to instantaneous stiffness with a damping factor of 0.05 at the initial
elastic stage.

8
Response under the 1985 Mexico Earthquake Motion

The earthquake resistant design before the 1985 Mexico Earthquake was governed
by the 1977 Construction Regulations for the Mexico D.F. The yield base shear
coefficient Fy was calculated for the lake bed zone (zone III):
T T
Fy  [ao  (c  ao ) ] /[1  (Q  1) ] for T  T1
T1 T1
Fy  c / Q for T1  T  T2
in which c=0.24, ao=0.06, r =1.0, T1=0.8 sec and T2=3.3 sec. Design ductility factors Q
was selected to be 1.0, 2.0 and 6.0 for the study. The yield resistance Fy was the same
for Q=4.0 and 6.0.
Three records (CDAF, CDAO, SCT1) were recorded in the lake bed zone in
Mexico City. These records showed long period contents in the waveform. The response
spectra of CDAF and SCT1 records exhibited large response at around 2.0 sec period,
whereas CDAO record showed a peak at around 1.3 to 1.5 sec. The maximum
acceleration of the EW component of SCT1 record was 1.68 m/sec2, and 0.70 to 1.00
m/sec2 in the other motions. Although the ground acceleration amplitudes were
relatively small, the response amplification was large by dominance in certain long-
period components.
The ductility demands (maximum response displacement divided by yield
displacement) are shown for Q=1.0 and Q=6.0 (or Q=4) in Figure 1.
(1) For a design ductility factor of Q=1.0, the systems did not yield against CDAF
and CDAO motions except at very short period. However, Record SCT1, especially the
EW component, caused yielding for systems at most period range. No structures should
have survived the EW motion of SCT1 unless much higher resistance was provided in
the system.
(2) For Q=2.0 (not shown in Figure 1), the ductility demand exceeded the target
design values under CDAF and CDAO motions for a period range longer than 1.3 sec.
Record SCT1 required ductility demand greater than the design value in all range of
periods.
(3) For Q=6.0 (or Q=4.0), the maximum response was comparable for the three
records. The ductility demand exceeded the design target for a period range less than 2.4
sec. The response increased as the system period decreased.
The nonlinear response analysis indicated that the systems designed with a small
ductility and large yield resistance developed small ductility demand, whereas the
systems designed with large ductility and small resistance developed a significantly
large plastic deformation.
A significant difference existed between the damage statistics and the calculated
ductility demand. If the structures were to possess the lateral resistance required by the
1977 building code, and if the structures were to fail at the design ductility Q, most of
the structures must have failed in the lake bed zone. On the contrary, the damage was
small in low-rise buildings, which indicates that the low-rise buildings must have been
designed for higher lateral resistance using lower design ductility factor Q. Furthermore,
the actual resistance of buildings is normally greater than the code required value
attributable to inherent additional resistance of the structural as well as non-structural
elements.

9
(a) Design Ductility Factor Q=1.0 (Elastic)

(b) Design Ductility Factor Q=4.0 or 6.0 (Very Ductile)


Figure 1: Response to the 1985 Mexico Earthquake in the lake bed zone

10
Response under the 1992 Erzincan Earthquake Motion

The governing design code was the 1975 Specifications for Structures in Disaster
Area (Part 1), the Ministry of Reconstruction and Settlement. The yield resistance Fy
was determined as follows:
Fy  Co K S I
where, Co: seismic zone coefficient (=0.10), K: structural type coefficient (=0.6, 1.0 and
1.5), S: spectral coefficient, and I: building importance coefficient (=1.0).

(a) North-south motion

(b) East-west motion


Figure 2: Response to the 1992 Erzincan Motion (K = 0.6, 1.0, 1.5)

11
The spectral coefficient S is given below:
1
S
0.8  T  To
in which To: soil natural period (=0.25 sec for hard soil), T: building natural period. The
soil was judged good from the micro-tremor measurement in the Erzincan.
A strong motion was recorded at the Meteorological Services Buildings, located in
the city approximately 5 km from the epicenter. Strong motion lasted approximately 20
sec with maximum acceleration of 0.5 g (g: acceleration of gravity) in the north-south
direction and 0.4 g in the east-west direction.
Ductility demand increased with a decreasing elastic period (Figure 2). For a six-
story building with structural type coefficient K of 1.0, ductility demand was
approximately 10 under the EW motion, and 25 under the NS motion; the ductility
demand was much beyond the deformation capacity of RC buildings. All RC buildings
must have collapsed if they were provided with resistance equal to the code required
value. However, most low-rise buildings, especially less than 3 stories, survived the
earthquake, indicating that the low-rise buildings in the area must have been provided
with lateral resistance much higher than the code specified value.

LESSONS LEARNED FROM RECENT EARTHQUAKES

Amenity of Life, Retrofitting and Building Law

The Ministry of Health and Welfare, Japan, investigated the causes of death in the
Kobe disaster on 5,488 cases; 4,816 (77 %) were killed by pressure and suffocation
under collapsed buildings. The fatal collapse of buildings was predominantly observed
in deteriorated traditional timber houses with heavy roofs. The heavy weight in a roof
was necessary for the heat insulation during hot summer days in Japan, and for the
protection of the roof from blow-up during annual typhoons (storms with strong winds).
The report also revealed that elderly people were affected worse than the young
people. The elderly could not afford to maintain the structure of their houses nor did
they wish to change their houses filled with past family memories and personal
attachments. Most of structural members in collapsed houses were eaten by termites or
rotten from weathering. The amenity and safety of daily living as well as the belief in
low probability of strong earthquake occurrence did not motivate the people to prepare
for a rare earthquake disaster. The government should have taken legal actions to update
the existing structure to the minimum safety level for the sake of public welfare.
Financial assistance should be considered to the elderly. In recent timber construction,
newly developed light heat insulation materials are used in the roof, and the roof is
attached rigidly to the structure to improve earthquake resistance.
The damage statistics demonstrated the poor performance of old buildings
designed using out-dated technology. The retrofit of deficient public buildings is an
urgent task of the government. The owner is responsible for maintaining his building to
the existing code level. The damage statistics clearly indicated that the number of
buildings suffering heavy damage or collapse was small even in the most heavily
damaged area. Therefore, an efficient and reliable screening procedure should be

12
employed to identify these probably deficient buildings from an entire existing building
stock. The upgrading of taller buildings should be emphasized.
The building law (or code) is a legal document that specifies a minimum design
level of structure of a building. Note that the law limits the constitutional freedom of
people for the sake of the public welfare. The legal requirements must be clear and
prescriptive so that the building official can easily check the conformance. The
minimum design level needs to be revised with the development of technology as well
as the change in socio-economic demand. Legal actions were necessary to update the
existing structure when the Building Law was revised. The owner of a building should
select the performance level higher than that required by the code.

Transfer of Technology from Research to Practice

Earthquake damage is caused by the lack of proper engineering practice. Some


examples are shown from the experience in Kobe and Erzincan.
(1) Construction materials and practices
A building must be constructed in accordance with design specifications about the
use of materials, the arrangement of reinforcement, the concrete work, etc. Poor quality
and placement of concrete was often observed in damaged buildings due to the lack of
technology. Poor concrete work is sometimes attributed to the error in structural design;
e.g., reinforcement may be congested in a section due to (a) the use of small cross
sectional area, (b) the use of lap splicing, and (c) the anchorage of beam reinforcement
in the already-congested beam-column joints. Construction work must be closely
inspected and guided by structural engineers during construction processes.
Construction workers must be trained as well as educated to handle right materials and
to execute the work properly.
(2) Importance of structural detailing in shear resistance
After the Kobe earthquake, the writer and his associates investigated the bend at
the end of hoop reinforcement in columns. Out of 64 heavily damaged or collapsed
buildings studied, forty-two buildings used 90-degree hooks at both hoop ends; twenty
buildings used combination of 90-degree hook at one end and 135-degree hook at the
other end. Only two buildings used either 135-degree hook at both hoop ends or welded
closed-shape hoops. Note that the 1933 Architectural Institute of Japan RC standard [4]
recommended the use of 135-degree hooks at the ends of a hoop. However, the
difficulty to bend a hoop end to 135 degree at the construction site prevented the worker
from using the 135-degree hook until two decades ago.
Columns of rectangular cross sections (e.g., roughly 200x500 mm or 300 x 600
mm) were often observed in Erzincan. These rectangular columns failed in shear in the
major principal direction (long dimension) because the amount of lateral reinforcement
did not meet shear input at flexural yielding in the major principal direction.
In Baguio City, Philippines, ties were closely spaced at the top and bottom of a
column probably to increase confining effect after flexural yielding. However, the
spacing and amount of shear reinforcement in the middle part of columns were not
sufficient to prevent shear failure before flexural yielding at the column ends.
The collapse of reinforced concrete buildings was often caused by brittle shear
failure of attributable to the use of (a) 90-degree hook at hoop ends, (b) wide spacing of
hoop reinforcement, (c) thin hoop reinforcement, and (d) pain bars as longitudinal
reinforcement.

13
(3) Beam-column connections
No lateral reinforcement was found within beam-column connections failing in
shear in Erzincan and Kobe. After spalling of concrete cover from the joint, the column
longitudinal reinforcement was observed to buckle.
The bottom reinforcement of a beam was often anchored straight in the beam-
column connection in Erzincan. This is a common detailing practice in non-seismic
regions, but will not allow flexural yielding under reversed loading. If a rectangular
column is used, the beam reinforcement cannot be anchored in the narrow side of the
beam-column connection.

Structural Planning

Structural engineers should consider the overall performance of a structure during


medium as well as high intensity earthquake motions;
(1) Resistance or ductility
A structure can be designed earthquake resistant (a) by providing large lateral
resistance but with limited ductility (strong structure) or (b) by providing large ductility
but with relatively small lateral resistance (ductile structure). Naturally, a ductile
structure suffers extensive structural as well as non-structural damages by a strong
earthquake, or suffers minor damages even by frequent medium-intensity earthquakes.
Consequently, the repair cost of the ductile structure must be significant. A strong
structure will not be damaged until the resistance is reached during an earthquake.
For this reason, the use of structural walls has been recommended for a long time.
Only 2 percent of box wall systems, normally designed following simple prescriptive
guidelines, suffered heavy damage (often in the foundation) or collapsed in Kobe [12].
(2) Weak-beam strong-column mechanism vs. Soft first-story mechanism
The damage of soft first-story buildings and the other buildings is compared in
Table 8 [12]. Less than three-quarters of the soft first-story buildings were operational
compared to ninety (90.2) percent of the other buildings. The collapse of soft first-story
structures was caused by brittle shear failure in the first story columns. The resistance
and ductility must be improved in this type of construction.

Table 8: Damage of soft first-story buildings during the Kobe Earthquake [12]
Operational Heavy
Structures Damage Damage Collapse Total
Soft First-story 275 54 51 380
Buildings (72.3) (14.2) (13.4) (100)
Other 3,186 174 171 3,531
Buildings (90.2) (4.9) (4.8) (100)

The weak-beam strong-column mechanism has been preferred by many structural


engineers because it dissipates earthquake-input energy at many distributed localities
and reduces ductility demand at each planned yield hinge region. However, it is
expensive to repair the damage at many scattered locations in the building after
earthquakes of high intensity as well as medium intensity. Contrary to common belief,
the damage in the soft first-story structure can be repaired using the present state of
construction technology as long as the first story does not collapse to the ground. A

14
structural engineer should consider the necessary repair cost in design.
(3) Non-structural elements and building content
For the life safety, the structure should naturally be prevented from collapsing. At
the same time, the response (acceleration or velocity) of a structure must be controlled
to prevent heavy furniture and equipment from overturning on the floor or to prevent
heavy equipment from falling from shelves; otherwise the contents of a building should
be properly fastened to the structure. The Ministry of Health and Welfare, Japan,
reported that 65 persons were killed under overturning heavy furniture.
Non-engineer residents were greatly scared by the damage of non-structural
elements, such as partitions, windows and doors. Stiff, weak and brittle brick walls,
filled in a flexible moment-resisting frame, fail at an early stage even during medium-
intensity earthquakes. Providing some gap on both side of a column could reduce such
damage. Non-structural elements must be protected from damage because the fall of
failed elements is dangerous for people escaping from the building, because the failed
elements may block doors from opening, and because the building may not be occupied
until the damaged elements are replaced. Furthermore, the cost of repair work is often
governed by the replacement of the damaged non-structural elements rather than the
repair work on structural elements.

Fire Prevention and Resistance

Large numbers of people were killed by fire in the 1906 San Francisco Earthquake,
U.S.A., and the 1923 Kanto (Tokyo) Earthquake. Fire is a major disaster after a strong
earthquake in a large urban area. In Kobe, the mortar or plaster cover protecting timber
construction fire fell off during a severe vibration, and the exposed timber structure
caught fire after the earthquake motion.

FUTURE OF EARTHQUAKE ENGINEERING

Contrary to our expectation, the damage statistics demonstrated that the percentage
of severely damaged buildings was extremely low for reinforced concrete and masonry
buildings even in the disastrous region. However, the impact of building collapse and
heavy damage to the society was quite large as evidenced by many mass media reports.
Building codes is intended to require minimum standard necessary for the life safety and
public welfare. Therefore, the building owner should select a design level higher than the
code specified level for the better function of the building and for the protection of
properties and investment.
A significant gap was observed between the expectations of structural performance
by the building owner (people) and the structural engineer. During the design stage, the
structural engineer should explain to the owner the expected performance of the
building under various intensity earthquake motions including the expected damage of
nonstructural elements. The cost of repair and occupancy after medium- and high-
intensity ground motions should be taken into consideration.
With the advice of the engineer, the owner should select the performance level of
his building suitable for its function and within the available funds, but the performance
level must be selected higher than the code minimum. The design engineer should
design a building to satisfy the owner's demand making best use of available materials

15
and structural engineering in the community. The methods to achieve the performance
level should not be limited, but the structural engineer should be able to choose most
suited procedure for a structure under design. The engineer should pay more attention to
safe design of tall buildings.
The construction engineer should construct the building in conformance with the
specifications of the structural engineer through close inspection and supervision by the
structural engineer.

REFERENCES

1. Housner, G.W. (1984), Historical View of Earthquake Engineering, Conference


Lecture, Proceedings, Eighth World Conference on Earthquake Engineering, San
Francisco, Post-Conference Volume, pp. 25 - 39.
2. Milne, J. and Burton, W.K. (1891) The Great Earthquake of Japan, 1891, Lane,
Crawford & Co., Yokohama, Japan.
3. Cross, H. (1930), Analysis of Continuous Frames by Distributing Fixed End
Moments, Proceedings, ASCE.
4. Architectural Institute of Japan (1933), AIJ Standard for Structural Calculation of
Reinforced Concrete Structures.
5. Biot, M.A. (1941) A Mechanical Analyzer for the Prediction of Earthquake Stresses,
Bulletin, Seismological Society of America, No. 31, pp.151-171.
6. Veletsos, A. S. and Newmark, N. M. (1960), Effect of Inelastic Behavior on the
Response of Simple Systems to Earthquake Motions, Proceedings, Second World
Conference on Earthquake Engineering, Tokyo-Kyoto, Volume II, pp. 895 - 912.
7. Seismological Committee (1959), Recommended Lateral Force Requirements,
SEAOC Code, Structural Engineering Association of California.
8. Applied Technology Council (1976), Tentative Provisions for the Development of
Seismic Regulations for Buildings, National Science Foundation and National
Bureau of Standards.
9. Architectural Institute of Japan (1987), Reports on the Damage Investigation of the
1985 Mexico Earthquake (in Japanese), 599 pp.
10. Architectural Institute of Japan (1992), Reports on the Damage Investigation of the
1990 Luzon Earthquake (in Japanese), 396 pp.
11. Architectural Institute of Japan (1993), Report on the Damage Investigation of the
1992 Turkey Earthquake (in Japanese), 221 pp.
12. Reinforced Concrete Committee (1996), Damage Investigation Report on Concrete
Buildings, the 1995 Hyogo-ken Nanbu Earthquake (in Japanese), The Kinki
Branch, Architectural Institute of Japan, 245 pp.
13. Takeda, T., Sozen, M. A., Nielsen, N. N. (1970), Reinforced Concrete Response to
Simulated Earthquakes, Journal, Structural Division, ASCE, Vol. 96, No. ST 12,
pp. 2557 - 2573.

16

Das könnte Ihnen auch gefallen