You are on page 1of 144

Previous | Next | Contents

ESDEP WG 17

SEISMIC DESIGN

Lecture 17.1: An Overall View of the

Seismic Behaviour of Structural Systems

OBJECTIVE/SCOPE

To give, by means of earthquake damage studies, an overall view of the principal forms of seismic
damage, together with some explanation of failure mechanisms.

PREREQUISITES

None.

RELATED LECTURES

None.

SUMMARY

It is not possible to design seismic resistant structures efficiently without understanding the ways in which
they are damaged by earthquakes in practice. The design process is not simply a matter of analysis,
calculation and following codes. A practical knowledge of building behaviour in earthquakes is essential.

Based on earthquake damage studies the principal forms of damage are described, together with some
explanation of the mechanics of failures.

1. PRESENTATION OF SLIDES OF EARTHQUAKE DAMAGE

The slides presented are not limited to structural steelwork buildings for two reasons. The first is that
many of the problems caused by earthquakes are common to all forms of construction. The second reason
is that it is very hard to find pictures of steel buildings which have suffered serious damage in an
earthquake!

2. DISCUSSION OF EARTHQUAKE DAMAGE

Engineers are generally accustomed to static loads acting on elastic structures. One of the most important
lessons learned from damage surveys is the difference in failure patterns between static loads applied in a
single direction and those due to cyclic loading. Another lesson is the necessity in earthquake design to
consider the behaviour of the main structural system after yielding.
SLIDE 1 The Pino Suarez building in Mexico City was a 21-storey structural steel framed building
constructed in about 1978. It suffered partial collapse and severe damage. Note the "K" bracings which
rely on the compressive strength of members - a system without the ductility to absorb damage without
collapsing.

An important aspect of post-earthquake study is the realisation of the important role that the quality of
construction plays. Earthquakes do not respect theories, calculations or divisions of responsibility. Many
instances of poor quality construction are invariably exposed in earthquake damaged buildings. Badly
placed reinforcement, poorly compacted concrete, incomplete grouting of masonry and loose or missing
bolts in structural steelwork are some of the commonest examples of poor quality.

Although the prime objective in the design of earthquake resistant buildings is the safety of the occupants
and passers by, every earthquake shows up numerous examples of lives at risk from minor faults in
construction - falling masonry or cladding, ceiling tiles dislodged, window frames separating from the
walls and toppling inwards or outwards, and escape paths blocked by jammed doors and fallen masonry.
Usually these types of failure could have been avoided with very little expense.

An important category of building failure in earthquake is the case where the building is so badly
damaged that it has to be demolished, although it has not collapsed. For the owner and the insurance
company the costs are similar whether the building collapses or is demolished. For the occupants it is the
difference between life and death.

Where two buildings are close together, or where there is a movement joint in a building, the two sides
are very likely to pound against each other during an earthquake. Major structural damage can result,
particularly where the floor levels differ. The cause lies in the closeness of the two structures and in the
flexibility of the buildings, factors which are within the control of the designer.
SLIDE 2 Adjacent buildings will pound against each other unless a sufficient space is allowed between
them. In this case in Mexico City the failure of a complete storey has been brought about by the two
buildings of differing height and dynamic properties pounding against each other.

Modern buildings are often assembled from many separate components. Older ones commonly have
timber floors with joists poorly tied to the supporting walls. Any lack of tying together in a building is
quickly exposed by seismic (earthquake) action. The nature of seismic ground motion inevitably leads to
differential movement between separate components, and in the absence of structural continuity,
differential movement will occur.

Aftershocks, generally of much smaller magnitude than the main seismic shock which they follow, play
no explicit part in the design process. Nevertheless they play a significant part in the immediate post-
earthquake rescue and survival operation. The further damage caused by aftershocks to already damaged
buildings is greater than their magnitude would otherwise suggest. Following major earthquakes many
structures brought to the brink of collapse by the main shock are destroyed by subsequent lesser shocks.

Concentrations of force occur where there are abrupt changes in structural stiffness or mass distribution.
For this reason building form should be regular and symmetrical as far as the functional requirements
permit.
SLIDE 3 Although the storey shear is normally greatest for the base storey, variations in the strength,
mass and stiffness can lead to failure initiating at any level - in this case a 'top down' failure in Mexico
City.

SLIDE 4 Finally some steelwork (undamaged) under construction in Los Angeles - an area of high
seismicity. This welded frame has to cope with the functional requirement of the building owner that
there should be fewer columns at the lowest storey for architectural reasons. This requirement is common
in hotels and office buildings where more open space is needed at this level.

3. GROUND BEHAVIOUR

The effects of violent shaking on the ground are to increase lateral and vertical forces temporarily, to
disturb the intergranular stability of non-cohesive soils, and to impose strains directly on surface material
where the fault plane reaches the surface. A transient increase in lateral and vertical forces places any soil
structures capable of movement at risk. The resulting types of damage are landslips and avalanches.
Experience of the 1970 earthquake in Peru and the 1964 earthquake in Anchorage, Alaska, show that this
damage may be on a massive scale. One village, Yungay, in Peru was destroyed almost entirely with the
loss of 18,000 lives by a debris flow involving tens of millions of tons of rock and ice.

The disturbance of the granular structure of soils by shaking leads to consolidation of both dry and
saturated material, due to the closer packing of grains. For saturated sands, pore pressure may be
increased by shaking to the point where it exceeds the confining soil pressure, resulting in temporary
liquefaction. This is an important effect. It can lead to massive foundation failure in bearing and piled
foundations, the collapse of slopes, embankments and dams. It can cause the phenomenon of "boiling"
where liquefied sand flows upwards in surface pockets. It is also possible for some unstable soils to
heave.

SLIDE 5 Serious ground failure may occur, especially in granular soils which are saturated. In this case
the soil failure caused the collapse of a number of dockside cranes at Vina del Mar in Chile.
SLIDE 6 This building in Mexico City has suffered an overturning foundation failure, most probably
initiated by failure of the supporting ground.

SLIDE 7 Settlement of foundations due to liquefaction or consolidation of the supporting soil can occur.
In this Mexico City building almost the whole of the ground floor storey has vanished below street level.
It is interesting to note that the building has not suffered a total collapse despite this settlement.

Shear movements in the ground may be at the surface or entirely below it. Where the earthquake fault
reaches the surface, permanent movements of considerable magnitude, in metres rather than centimetres,
may occur. Surface shear movements may also take place as a result of other soil displacement - landslips
or consolidation for example. Sub-surface shear movements can occur in weaker strata, leading to damage
of embedded or buried structures. Sub-surface shear movements also reduce the transmission of ground
motion to the surface, effectively putting an upper bound on the surface motion.

In considering the more spectacular permanent ground displacements that can result from ground shaking,
it should not be forgotten that elastic displacements also occur. They are critical in the design of piles,
underground pipelines and culvert-type structures. Failures in underground piping and ductwork are
common in earthquakes and have important implications for the post-earthquake emergency services.

4. SOIL STRUCTURE INTERACTION

The nature of ground shaking is substantially influenced by the subsoil underlying the site. Soft soils tend
to vibrate at a lower frequency than hard ground sites but are likely to have a higher peak acceleration.
General indications of the effect of soil on the fundamental period of surface motion can be seen from
Figure 1. The importance of the period in practice is the increased liability of damage where the natural
period of the building is close to that of the ground. For low amplitude shaking, quite large amplifications
are possible. In very soft soil, for example, amplifications of over 20 have been recorded for San
Francisco Bay mud. However this effect is swiftly overcome by yielding of soft soils as amplitudes
increase, so that, for strong shaking, peak accelerations are normally reduced in transmission through the
upper levels of soil.

Considering the subsoil layers as a dynamic system, it is clear that surface responses will be modified if
another structure is added at the upper level. The interaction of the structure and its supporting soil falls
into two categories. Buildings in general are light in relation to the mass of the supporting soil and
relatively flexible. Thus the addition of the building does not affect the surface ground motion
significantly. However local flexibility of the soil where it is in contact with the foundation can modify
the building response. The effects of this local flexibility are to modify vibration modes, lower natural
frequencies and generate additional damping through energy dissipation in the surrounding soil. Although
an increase in response can occur, the general effect is to produce a reduction in base shear. Piled
foundations, in comparison with bearing foundations, generally have a lesser effect on the mode shapes
and frequencies but produce lower damping effects.

The second type of soil structure interaction to be considered is where a structure is massive and rigid. In
this case the structure becomes a significant element in the dynamic system represented by the subsoil and
the structure. It causes the surface ground motion in its vicinity to be modified.

5. THE BEHAVIOUR OF FOUNDATIONS

Failures of building foundations in earthquakes are not uncommon but are nearly always caused by failure
of the supporting soil. Overturning failures due to uplift occur rarely, far less often than calculations
would suggest. This rarity is probably due to the effective reduction in stiffness that accompanies uplift,
which correspondingly reduces the force exerted by the ground acceleration. There can be no doubt that
substantial tension from overturning forces can develop at foundation level. Examination of some lower
failed columns in Caracas, following the 1967 earthquake, showed that they had failed in tension due to a
combination of overturning forces and vertical ground acceleration.

Instances of failure in piles have been reported. In general, piles tend to conform with ground
displacements and are vulnerable at points where adjacent strata have markedly different properties. Some
configurations incorporating raked piles have failed at the underside of the pile cap.

6. THE RESPONSE OF STEEL FRAMED STRUCTURES

Generally steel framed structures are engineered structures competent to resist gravity and wind loads. In
the familiar processes of design, attention is commonly given to stresses before considering displacement.
The secondary effects of displacement are often forgotten. Earthquake damage frequently draws attention
back both to the direct effects of large displacements, such as the pounding at joints and damage to non-
structural components and contents, and to the second order effects caused by large displacements.

Buildings with shear walls or braced frames, as long as they maintain their integrity, compare favourably
in performance with more flexible framed structures as far a damage to contents and non-structural items
is concerned. Particular points commonly revealed for framed structures are:

i. Corner columns often behave badly in comparison with other exterior and interior columns. This
behaviour suggests that the effects of earthquake forces in orthogonal directions are not adequately dealt
with in design.

ii. Complete failure in members detailed for ductility is rare. Where members with low ductility have
failed it is clear that failure is swift. This behaviour is particularly marked in reinforced concrete
members.

iii. The maximum practicable redundancy is shown to be desirable. The failure mechanism should involve
as many members as possible, providing alternative load paths when one member yields or fails.

iv. Where yielding occurs in the columns before the beams, failure of the frame becomes much more
likely. This point is illustrated in Figure 2 which shows the number of ductile hinges needed for failure in
the column yielding mode compared with the beam yielding mode.
Structural steelwork shows the following types of damage from earthquakes:

i. Brittle failure of bolts in shear or tension.

ii. Brittle failure of welds, particularly fillet welds, in shear or tension.

iii. Member buckling, including torsional buckling.

iv. Local web and flange buckling.

v. Uplift of braced frames.

v. Local failure of connection elements such as cleats and tees.

vii. Bolt slip.

viii. High deflections in unbraced frames.

ix. Failure of connections between steel members and other building elements, such as floors.

x. Anchorages of components into masonry or concrete by cast in or expanding head bolts are almost
invariably brittle in shear and tension. Thus they are unable to accommodate any movement. Accordingly
failures are commonplace, aggravated when the masonry or concrete into which the anchorage is placed is
also damaged.

xi. Many failures occur in horizontal torsion, especially in structures where the centres of mass and
resistance are some distance apart on plan, or where the inherent torsional resistance of the system is low.
A common case of a torsionally vulnerable structure is where buildings are located at street corners.

SLIDE 8 This Mexico City building experienced failure of ground floor columns due to a soft first storey
and horizontal torsional effects.

7. THE BEHAVIOUR OF FLOORS

Floor slabs function as diaphragms in transferring lateral forces. Figure3 shows two possible floor plans.
In the first case there is very little diaphragm action but, in the second it is clearly significant. The transfer
of shear at each end wall imposed high stresses in the slab. Some fully or partially prefabricated floor
systems have very little strength in horizontal shear or bending.
SLIDE 9 Horizontal diaphragms are not always rigid elements capable of distributing forces between
frames. In this Anchorage school a reinforced concrete roof slab has torn like a piece of cardboard.

8. THE BEHAVIOUR OF SECONDARY STRUCTURES AND APPENDAGES

Appendages to buildings - masonry parapets, penthouses, roof tanks, cladding and cantilevers - tend to
behave badly in earthquakes. The reasons for this are twofold. Firstly many of them are designed without
any ductility, and secondly the effects of dynamic amplification by the building to which they are
attached may greatly increase the forces applied to them.

Figure 4 illustrates the effect of the dynamic response of the building on the response spectrum,
comparing the ground level spectrum with that at the fifth floor. The peak values are both amplified and
shifted in frequency.
The contents of buildings often suffer major damage even when the building itself is relatively unharmed.
This effect is greater for more flexible buildings. It represents an additional reason for the designer to
exercise close control over displacements. In many modern buildings the contents are of greater value and
importance than the building itself. The costs of preventing damage are often trivial, for example, use of
steel angle ties to the tops of racks and floor bolts to shelving.

At any level in a multi-storey building the ground motion will be modified by the motion of the building
itself. Generally the effect is to concentrate the frequency of response around a band close to the natural
frequency of the building, and to amplify the peak acceleration roughly in proportion to the height,
reaching an amplification of perhaps two or three at roof level.

Any contents which are either very stiff or which have a natural frequency of their own close to that of the
building are therefore subjected to greater forces than they would experience if mounted at ground level.
Experience shows that non-structural items which are suspended such as ceiling systems and light fittings
perform badly. Appendages such as parapets and mechanical plant also suffer high levels of damage,
especially where they function as single degree of freedom "inverted pendulums". Damage also increases
in multi-storey structures towards the roof. Roof tanks and penthouses are also subject to high forces.

SLIDE 10 All these cladding panels have fallen during the earthquake in Vino del Mar, Chile, creating a
serious hazard for any occupants running for safety from the building. Cladding needs to be attached with
ductile fixings capable of substantial deformation without fracture.

SLIDE 11 These batteries formed part of the emergency power system in a California hospital in 1972.
During an earthquake the batteries fell off their racks and did not function when they were needed.
Patients on life support systems died as a result. The contents of buildings are often of great value or
importance and can be protected by limiting displacements and by simple cheap measures. In this case the
batteries could have been strapped down or clipped to the racks which should have been bolted to the
floor.
SLIDE 12 Surface finishes also present a major hazard when they fall, as in the case of this Mexico City
building.

SLIDE 13 Experience with appendages to buildings such as this Mexico City water tank are that they
perform badly in earthquakes. Dynamic response analysis also supports this experience. There is in effect
a major discontinuity at the junction of the building and the tank with a resultant high stress
concentration.
SLIDE 14 This Mexico City building illustrates the fragility of curtain wall glazing systems. They were
unable to cope with the differential movement of the floors to which they were attached.

9. THE BEHAVIOUR OF MASONRY AND CLADDING

Failure of unreinforced masonry is so common that it is almost taken for granted and forgotten. Many
earthquake codes ban the use of unreinforced masonry altogether. However, economic reasons still ensure
that it is very widely used both for low-rise structural walls and as infill to framed structures.

Failures of both reinforced and unreinforced masonry in-plane are common. Masonry is very stiff and
brittle in-plane so that the forces transmitted by ground shaking are high and failure is accompanied by a
marked reduction in strength and stiffness. Damage normally comprises either collapse or diagonal
cracking in both directions ("X" cracking). Cracks will often be concentrated around openings. Cracking
will frequently follow the mortar joints.
SLIDE 15 Typical `X' cracking of masonry in this Anchorage, Alaska school illustrates the effect of
reversing horizontal shear forces during the earthquake. Shear stresses are concentrated opposite the
window openings.

SLIDE 16 Where masonry buts up against a structural column it has the effect of concentrating shear
over a short length so that the member can fail in shear (brittle failure) instead of bending (ductile failure).
This behaviour is generally referred to as the short column effect.

The full implications of frame-infill masonry behaviour are complex. The failure of walls out-of-plane is
common and causes substantial secondary damage.

Figures 5 to 7 illustrate the interaction of infill masonry and frame in the in-plane direction. Figure 5
shows the interaction of the undamaged masonry panel with the frame. The masonry acts as a diagonal
compression brace in the direction of the arrow, resulting in a substantial stiffening of the frame and
redistribution of bending moments and shears in the frame. Figure 6 shows the effect of the horizontally
sheared panel and accompanying rearrangement of the frame forces. Once the panel has sheared the effect
of the diagonal compression zone is lost. Figure 7 shows the situation where the masonry does not occupy
the whole of the panel, resulting in high shear forces in the unsupported portion of the column.
The redistribution in plan of forces, due to the stiffening effect of the infill masonry, is also of
consequence. The frame may be stiffened leading to higher dynamic forces and accidental eccentricity
leading to high torsional forces may result.

Some elements can be damaged by drift, or inter-storey displacement. Windows and cladding elements
are frequently connected rigidly to more than one level and, if there is no ductile provision for relative
movement in the connections, they may fail.

10. TANKS

Steel tank structures are a specialised area dealt with in Lecture 17.6. They suffer from compression
failure in the tank wall (including "elephant's foot" buckling) and tearing of the wall-floor joint.

11. CONCLUDING SUMMARY

 Failure patterns resulting from static loads applied in a single direction differ from those due to
seismic loading.
 Adjacent structures may pound against each other unless a sufficient space is allowed between
them.
 The behaviour of the main structural system after yielding must be considered.
 Poor design and poor quality construction are invariably exposed in an earthquake.
 Minor faults in construction can create risk to life - falling masonry, windows, etc.
 Soil structure interaction plays on important role.
 Constructions with shear walls or braced frames "perform" favourably.

12. ADDITIONAL READING


1. Dowrick, D. J., "Earthquake Resistant Design", John Wiley Second Edition 1989.
2. Key, D. E., "Earthquake Design Practice for Buildings", Thomas Telford 1988.
3. Naeim, F., "Seismic Design Handbook", Van Nostrand Rheinhold 1989.
4. "Earthquake Spectra", Earthquake Engineering Research Institute, 6431 Fairmount Avenue, Suite
7, El Cerrito, California CA94530, USA.
5. "Earthquake Engineering & Structural Dynamics", John Wiley.

Previous | Next | Contents

Previous | Next | Contents

ESDEP WG 17

SEISMIC DESIGN

Lecture 17.2: Introduction to Seismic

Design - Seismic Hazard and Seismic

Risk

OBJECTIVE/SCOPE

To give an introduction to seismicity, seismic hazard, seismic risk, and seismic measures.

PREREQUISITES

None.

RELATED LECTURES

None.

SUMMARY

The lecture introduces seismicity, explaining the origins of earthquakes and summarises their
characteristics in both general and engineering terms. The need for probabilistic assessments is
demonstrated and the concept of response spectra is introduced. The basic approaches for design against
earthquakes and Eurocode 8[1] are presented.

1. INTRODUCTION

Among the natural phenomena that have worried human kind, earthquakes are without doubt the most
distressing one. The fact that, so far, the occurrence of earthquakes has been unpredictable, makes them
especially feared by the common citizen, for he feels there is no way to assure an effective preparedness.

The most feared effects of earthquakes are collapses of constructions, for they not only usually imply
human casualties but represent huge losses for individuals as well as for the community. Thus, although
other consequences of earthquakes may include landslides, soil liquefaction and tsunamis, it is the aim in
this lecture to study seismic motion from the point of view of the natural hazard it poses to construction,
and particularly to steel structures.

The fundamental goals of any structural design are safety, serviceability and economy. Achieving these
goals for design in seismic regions is especially important and difficult. Uncertainty and unpredictability
of when, where and how a seismic event will strike a community increases the overall difficulty. In
addition, lack of understanding and ability to estimate the performance of constructed facilities makes it
difficult to achieve the above mentioned goals.

The future occurrence of earthquakes can be regarded as a seismic hazard, whose consequences represent
what can be defined as seismic risk. The separate study of these two concepts is important. The first
represents the action of nature and the second the effects on mankind and man-made structures.

2. THE SEISMIC EVENT

2.1 General

The knowledge and study of past seismic events is an important way of predicting the potential seismic
hazard for the different zones of the earth. Earthquakes have been reported as far back as during the
Babylonian Empire or in 780 BC in China.

A region which has suffered large earthquakes (Figure 1) is the circum-Pacific belt including New
Zealand, the Tonga and New-Hebrides Archipelagos, the Philippines, Taiwan, Japan, the Kurile and
Aleutian Isles, Alaska, the western coasts of Canada and the United States, Mexico, all the countries in
Central America and the western coast of South America from Colombia to Chile. Other regions of the
world that also have been subject to devastating earthquakes in the past are the northern and eastern zones
of China, northern India, Iran, the south of the Arabian Peninsula, Turkey, all the southern part of Europe
including Greece, Yugoslavia, Italy and Portugal, the north of Africa and some of the Caribbean
countries.
Worldwide, the most devastating seismic event which has ever happened is believed to be the 1556,
January 23rd earthquake in the Shaanxi Province of China. That earthquake may have caused more than
half a million casualties. More recently, two other Chinese provinces, the Ningxia province in 1920 and
the Hebei province in 1976, were hit by earthquakes that may have caused several hundreds of thousands
of dead.

In Europe, earthquakes are reported as far back as 373 BC in Helice, Greece. Other catastrophic
earthquakes in Europe occurred in 365, 1455 and 1626 in Naples, 1531 and 1755 in Portugal, 1693 in
Sicily, 1783 in Calabria and 1908 in Messina. Each one of these earthquakes is believed to have caused
between 30000 and 60000 deaths. Even if these figures are not totally reliable, they give a dimension of
the consequences or the risk that may result from the seismic hazard in some European countries.

These major earthquakes have each caused not only a large number of human casualties due to the
collapse of houses and other buildings, but also have caused huge economical losses which in some cases
took long periods to recover. The large losses, human and economic, that can be expected from the
occurrence of future earthquakes justify special attention being given to the study of earthquake
phenomena and the earthquake hazard.
2.2 Origins of Earthquakes

Earthquakes have their origin in the sudden release of accumulated energy in some zones of the earth's
crust and the resulting propagation of seismic waves.

Wegener introduced the concept of continental drift to explain the origin of the continents, and why the
earth's crust is divided into interacting plates. The zones of the earth where most earthquakes are
generated are at the boundaries of the plates. Earthquakes occur in some cases due to subduction
movements between two plates, as is the case of the Pacific plate which moves underneath the South
American continent, and in other cases due to sliding movements between the two plates, as is the case of
San Andreas fault in California. In Southern Europe the boundary between the African and the
Euroasiatic plates is responsible for some very large earthquakes, as for example the 1755 earthquake that
destroyed most of the city of Lisbon.

Other zones where earthquakes occur are at the faults in the intraplate regions, due to the accumulation of
strains caused by the pressures in the plate's boundaries. Most of the Chinese earthquakes are generated in
the intraplate region. In Europe a similar region is involved for most of the southern part of the continent
but also for some other central and northern areas.

The point or the zone at which the earthquake slip first occurs is commonly designated as the focus or
hypocentre. The earthquake focus is usually at a certain depth, known as the focal depth. The intersection
of a vertical line through the focus with the ground surface is known as the epicentre (Figure 2).
Obviously the most affected zones are the ones closer to the focus, showing that distance to the epicentre
(or hypocentre) is a significant factor of seismic hazard.
The sudden release of energy at the focus generates seismic waves that propagate through the rock and
soil layers. There are three basic types of seismic waves; P waves, S waves and surface waves which
include the Love and Rayleigh waves. The difference of velocity between the P and the S waves allows,
by means of the difference in the arrival time, the determination of the hypocentral distance. Typical
velocities of P and S waves vary from 100m/sec for S waves in unconsolidated soils (300m/sec for P
waves) to 4000m/sec for S waves in igneous rocks (7500m/sec for P waves).

2.3 Earthquake Characteristic

The "size" of the earthquake or what could be seen as a seismic scale is a very important factor for a
correct characterization of its potential hazard. Intensity and magnitude are two different means of
"measuring" an earthquake which are often confused by the media.

The concept of magnitude which was first introduced by Richter and which still carries his name,
represents a measure of the earthquake that is supposed to be independent of the location at which the
measurement is obtained. It is related to the amplitude of the seismic waves corrected with respect to
distance. It represents a universal measure of the size of the earthquake, independently of its effects.
Although there is no maximum value for the magnitude of an earthquake, the two largest magnitudes ever
observed correspond to the 1906 earthquake off the coast of Ecuador and the 1933 earthquake off the
Sanriku coast in Japan with magnitudes of 8,9. The 1755 earthquake, off the coast of Portugal, is believed
to have been the largest earthquake in Europe with a magnitude of 8,6.

The magnitude of an earthquake can be related to other physical measures of earthquakes such as the total
released energy, the length of the fault rupture, the fault rupture area and the fault slippage or relative
displacement suffered between the two sides of the fault. Several relationships have been proposed by
different authors. The ones presented here are merely an indication of the types of relationships. More
accurate expressions can probably be presented for different seismic zones. Approximate relationships
between magnitude (M), total energy (E in ergs), fault rupture length (L in meters), fault rupture area (A
in Km2) and fault slippage displacement (D in meters) are:

Log E = 9,9 + 1,9 M - 0,024 M2

M = 1,61 + 1,182 log L

M = 4,15 + log A

M = 6,75 + 1,197 log D

The relationship between energy and magnitude shows that an earthquake of magnitude 8 releases as
much as about 37 times the energy released by a magnitude 7 earthquake. The same observation can be
made for the relationships between magnitude and measures of the fault, showing that an increase of one
degree in the Richter scale corresponds to a considerable increase in terms of seismic hazard.

A different way of measuring an earthquake, has been adopted, based on a scale initially proposed by
Mercalli and later modified, known as the Modified Mercalli Intensity (MMI). According to this scale
(Table 1), which varies between I and XII, the intensity of an earthquake is dependent on the observed
effects on landscape, structures and people at a given site. Thus, the intensity is variable from place to
place and relies on a subjective appreciation of the earthquake consequences. An approximate
correspondence between MMI and ground acceleration, a parameter which will be further discussed, is
presented in Table 1.

Table 1 Modified Mercalli Intensity (MMI) Scale Peak ground


acceleration

(m sec-2)
I Not felt by people. < 2,5 x 10-3
II Felt only be a few persons at rest, especially on upper floors of buildings. 2,5 x 10-3 - 0,005
III Felt indoors by many people. Feels like the vibration of a light truck passing
by. Hanging objects swing. May not be recognised as an earthquake.
IV Felt indoors by most people and outdoors by a few. Feels like the vibration of 0,005 - 0,010
a heavy truck passing by. Hanging objects swing noticeably. Standing
automobiles rock. Windows, dishes, and doors rattle; glasses and crockery
clink. Some wood walls and frames creak.
V Felt by most people indoors and outdoors; sleepers awaken. Liquids disturbed,
with some spillage. Small objects displaced or upset; some dishes and 0,010 - 0,025
glassware broken. Doors swing; pendulum clocks may stop. Trees and poles
may shake.

0,025 - 0,05
VI Felt by everyone. Many people are frightened; some run outdoors. People 0,05 - 0,10
move unsteadily. Dishes, glassware, and some windows break. Small objects
fall off shelves; pictures fall off walls. Furniture may move. Weak plaster and
masonry D cracks. Church and school bells ring. Trees and bushes shake
visibly.
VII People are frightened; it is difficult to stand. Automobile drivers notice the 0,10 - 0,25
shaking. Hanging objects quiver. Furniture breaks. Weak chimneys break.
Loose bricks, stones, tiles, corners, unbraced parapets, and architectural
ornaments fall from buildings. Damage to masonry D; some cracks in
masonry C. Waves seen on ponds. Small slides along sand or gravel banks.
Large bells ring. Concrete irrigation ditches damaged.
VIII General fright; signs of panic. Steering of vehicles is affected. Stucco falls; 0,25 - 0,5
some masonry walls fall. Some twisting and falling of chimneys, factory
stacks, monuments, towers, and elevated tanks. Frame houses move on
foundations if not bolted down. Heavy damage to masonry D; damage and
partial collapse on masonry C. Some damage to masonry B, none to masonry
A. Decayed piles break off. Branches break from trees. Flow or temperature of
water in springs and wells may change. Cracks appear in wet ground and on
steep slopes.
IX General panic. Damage to well-built structures; much interior damage. Frame 0,5 - 1,0
structures are racked and, if not bolted down, shift off foundations. Masonry D
destroyed; heavy damage to masonry C, sometime with complete collapse;
masonry B seriously damaged. Damage to foundations, serious damage to
reservoirs; underground pipes broken. Conspicuous cracks in the ground. In
alluvial soil, sand and mud is ejected; earthquake fountains occur and sand
craters are formed.
X Most masonry and frame structures destroyed with their foundations. Some 1,0 - 2,5
well-built wooden structures and bridges destroyed. Serous damage to dams,
dikes, and embankments. Large landslides. Water is thrown on banks of
canals, rivers, and lakes. Sand and mud are shifted horizontally on beaches
and flat land. Rails bent slightly.
XI Most masonry and wood structures collapse. Some bridges destroyed. Large 2,5 - 5,0
fissures appear in the ground. Underground pipelines completely out of
service. Rails badly bent.
XII Damage is total. Large rock masses are displaced. Waves are seen on the 5,0 - 10,0
surface of the ground. Lines of sight and level are distorted. Objects are
thrown into the air.
1. At intensity I there may be effects from very large earthquakes at considerable distance in the
form of long-period motion. These effects include disturbed birds and animals, swaying of
hanging objects, and slow swinging of doors, although people will not feel the shaking and will
not recognize the effects as being caused by an earthquake.
2. Each earthquake effect is listed in the table at the level of intensity at which it appears
frequently. It may be found less frequently or less strongly at the preceding (lower) level and
more frequently and more strongly at higher levels.
3. The quality of masonry or brick construction was categorized by Richter (1956) as follows:

Masonry A Good workmanship, mortar, and design; reinforced, especially laterally, and bound together
by using steel, concrete, etc: designed to resist lateral forces.

Masonry B Good workmanship and mortar; reinforced, but not designed in detail to resist lateral forces.

Masonry C Ordinary workmanship and mortar; no extreme weaknesses like failing to tie in at corners,
but neither reinforced nor designed against horizontal forces.

Masonry D Weak materials, such as adobe; poor mortar; low standards of workmanship; weak
horizontally.

Figure 3 represents a map of the maximum observed intensities in Europe, which is based on the
recollection of the effects of past earthquakes, and thus can already be looked at as a measure of seismic
risk.
The duration of ground motion is another parameter of great interest when assessing the seismic hazard
for a given seismic environment. Although there is no single definition for the duration of an earthquake,
all the most commonly used definitions agree as a rule that the duration of an earthquake at a given site
increases with the magnitude, epicentral distance, and the depth of the soil above bed-rock. The duration
of an earthquake is a very important parameter especially when assessing the non-linear response of
structures. The accumulation of structural damage, which is related to the non-linear behaviour of the
structure and may lead to structural failure, can be highly affected by the total time a structure is subjected
to strong ground motion. An earthquake with a given magnitude may represent a smaller hazard than an
earthquake with a smaller magnitude but larger duration or even than a series of smaller magnitude
earthquakes.
All the possible measures of an earthquake that have been presented so far, are of limited interest from the
engineering point of view. The relationships that have been established between the different parameters
are not deterministic and involve a great amount of uncertainty and variability. On the other hand, they
relate more to the physical aspects of the seismic source and, except for the Mercalli Intensity which is
determined based on subjective judgement, do not take into account the site characteristics and the
hypocentral or epicentral distance.

The need for an engineering characterization of the seismic motion, justifies the use of alternative
parameters, such as the maximum ground acceleration or peak ground acceleration (ag) observed during
the ground motion at a given site. The maximum acceleration has been observed to be statistically
dependent on the magnitude of the earthquake. Hence it is dependent on the severity of the seismic
source, and is also highly dependent on the distance to the epicentre and on the soil characteristics and
other local site conditions. Figure 4 shows the type of relationship that exists between ag and distance for
different earthquake magnitudes.

Approximate relationships exist between the Richter Magnitude, the Modified Mercalli Intensity and
ag observed in the epicentral zone. However these relationships are very dependent on several other
parameters such as the local soil conditions and even on the type of seismic source.

Instruments are available that measure the movements of the ground due to earthquakes. Some
instruments measure the ground displacements and are called seismographs. To measure the ground
accelerations, other type of device exist, called accelerographs. The accelerographs register the
accelerations of the soil and the record obtained is called an accelerogram. A typical accelerogram is
represented in Figure 5, showing the peak ground acceleration (ag).

Knowing, for a given earthquake and site, the accelerations in three orthogonal directions, it is possible to
evaluate the response of a structure when subjected to that specific earthquake.

But for a given site, there may be more than one potential seismic source and from a given source
earthquakes with different magnitudes, durations and peak ground accelerations may occur. In addition,
even for the same earthquake, accelerograms obtained in different locations may vary substantially,
depending on the local site conditions. The geometry and properties of the soil have been shown in past
earthquakes to have a large influence on the characteristics of the accelerogams obtained. Thus, the
accelerograms obtained from past earthquakes have to be used with special care. They may not correctly
represent the ground accelerations of future events.

The knowledge of the seismic ground motion is an essential aspect of the characterization of the seismic
hazard. Access to accelerograms from different earthquakes, in different seismic environments, for
several magnitudes and epicentral distances, in different soil conditions gives a unique basis for
characterizing the ground motion and determining its most influential parameter. Arrays of strong ground
motion accelerographs have been used in the last decade allowing a more reliable estimate of the
earthquake motion. Thus a probabilistic assessment of the earthquake input is obtained for use in
engineering applications.

Among the aspects that are investigated with arrays of ground motion accelerographs are the influence of
the type of seismic action, hypocentral distance, wave propagation path, orientation of the site with
respect to the fault line, local soil conditions and local topography.

During the lifetime of a structure there is a certain probability that it may be subjected to one or more
earthquakes. The probability depends on the seismic environment and on the period for which the
structure is to function. The probability that an earthquake with a large magnitude, and consequently with
large ag values, occurs during the lifetime of a structure, is smaller than the probability of occurrence of
smaller earthquakes. The number of earthquakes (N) having a magnitude (M) or greater per year, can be
estimated by means of recurrence formulae of the type.
log N = a - b M

where a and b are parameters depending on local conditions.

For each seismic zone, and based on past seismic events, recurrence formulae can be obtained, giving the
annual probability of occurrence of earthquakes with a certain magnitude, or the return period of
occurrence of an earthquake with a given magnitude. As the magnitude can be related with ag, these types
of relationship give the return period of occurrence of a certain level of ground acceleration. According to
the time period to be adopted, which depends on the level of risk to be accepted, the corresponding
ag value can be determined. This ag value, is the peak ground acceleration that will be exceeded with a
given probability, necessarily very small, and thus assuming a certain level of seismic risk.

Differences between past and future ground accelerations will exist not only in terms of the maximum
observed values (ag) but also in terms of the frequency content. Thus, another aspect that has to be
examined in any study of seismic hazard, is the frequency content of the earthquake records. The fourier
transform, the spectral density function or power spectrum and the response spectrum are different ways
to characterize an accelerogram in the frequency domain. It should be noted that Eurocode 8
recommendations allow the use of accelerograms, power spectra or response spectra to define the seismic
motion for structural analysis purposes. The last approach will be discussed here because it is the simplest
approach of those available which have direct application to structural analysis.

2.4 Response Spectrum

The response spectrum of a given earthquake record is the representation of some maximum response
quantity of a damped, linear, single degree of freedom system as a function of the natural frequency of
that system.

For example, for the system shown in Figure 6, with mass m, stiffness k, (velocity dependent) damping c,
ground displacement dg, and displacement of the mass relative to the ground dr, the equation of motion
can be written in the form

m (dg'' + dr'') + cdr'' + kdr = 0


or

mdr'' + cdr' + kdr = - mdg''

This equation of relative displacement is the same as that for a mass with fixed base subjected to a
horizontal force -mdg''. Introduction of the natural frequency of the undamped system  = , the
natural period of the undamped system T = 2/, and the damping ratio  = c/2m, gives

dr'' + 2 dr' + 2 dr = -dg''

with the solution

dr = -exp (-p,t)/D dg''() exp [ D (t - )] sin D (t - ) d ,

di = exp (- D t)/D dg () exp [ D (t - )] sin D (t - ) d ,

where

D = (1 - 2) is the natural frequency of the damped system.

 =1 corresponds to the critical damping ccr = 2 .

For a given accelerogram, i.e. given dg'', the maximum of dr, for a given value of , can be determined for
each D. Usually the value  = 0,05 is used as a reference value and a correction factor  for damping
ratios different from 5% is introduced.

A typical acceleration response spectrum for three damping ratio values is shown in Figure 7.
The two parameters that most influence the shape of the response spectrum, or its frequency content, are
the type of earthquake and the local soil conditions. The influence of these two parameters on the shape of
the spectrum arises from the phenomenon of resonance. In reality the fact that a certain earthquake has a
predominance of energy centred in a given frequency range will cause the response spectrum to have
larger amplitudes in that same frequency range. Two aspects that may lead to different spectra are the
distance of the site to the seismic source and the characteristics of the local soil. Large hypocentral
distances tend to diminish the high frequency components of the local ground motion. Soft soils also tend
to amplify the low frequency components of the ground motion, whereas for hard soils the high frequency
components are amplified.

In the past, it has been observed that similar structures subjected to the same earthquakes show a quite
different seismic behaviour because of the local soil conditions. In the 1967 Caracas, Venezuela
earthquake, it was observed that damage to buildings was not uniform throughout the city. Tall buildings
with foundations on soft thick soil layers showed much more damage than the same type of building with
foundations on stiffer soils. The opposite was observed for low-rise buildings; they showed more damage
for foundations on the stiffer soils. This observation showed that the same earthquake motion can be
filtered in a different way by two distinct soils. Thus the seismic input into a structure may vary according
to the local soil conditions. The interaction between the ground motion and the structural characteristics is
thus of great importance in the evaluation of the seismic response of structures and the associated seismic
risk.

3. EARTHQUAKE INPUT FOR STRUCTURAL DESIGN

The fact that, for a given earthquake source and site, there have been no observed earthquakes with a
magnitude, intensity, or peak ground acceleration larger than certain values, does not mean that larger
values will not be observed in future. Thus, the maximum possible or probable values have to be derived
using a probabilistic approach. Furthermore, if one derives probabilistic maximum values for earthquakes
that may occur during a certain future period of time, the values will differ from the ones relating to a
different period of time. The return period of an earthquake with given characteristics, can be defined as
the inverse of the annual probability of occurrence of that event. The larger the seismic event, the larger
the corresponding return period as shown by the recurrence formulae already presented.

If the earthquake for which the structure has to be designed and its return period are known, and if the
period for which the structure is designed is also known, the probability of the structure being subjected to
the earthquake during its lifetime can be determined. Evaluating this probability is a matter of assessing a
parameter of seismic risk. To evaluate the global seismic risk, one should combine this type of
information with the information regarding the single probability of collapse or malfunctioning of the
structure if designed according to certain levels and standards of resistance and ductility.

Different earthquakes lead to dissimilar response spectra. Not only different maximum values of the
ground acceleration (ag) lead to different maximum spectrum values, but also different accelerograms will
result in dissimilar shapes of spectra even with the same ag. So, the use of response spectra to characterize
a certain potential seismic event, has to take into account the influence of important aspects such as the
nature and distance of the seismic source and the characteristics of the soil.

For these reasons, the evaluation of response spectra for design purposes must include a probabilistic
study of the seismic occurrences. The study will define the maximum ground acceleration and the shape
of the spectra to be considered, for each seismic source and each different kind of soil. This definition is
usually obtained by statistical means. The spectra used for design purposes, and the spectra presented in
regulations are usually the smoothed graphs of the maximum credible values of the corresponding
spectra, for a certain level of risk acceptance, in terms of seismic origin and local soil conditions, obtained
for different earthquakes.

The different levels of risk acceptance are also related to the importance of the structure to be designed.
The catastrophic consequences arising as a result of collapse or malfunctioning of important buildings and
other structures, such as hospitals, fire stations, power plants, schools, dams, main bridges, etc. requires
design to a lower level of risk than for normal structures. This lower level is achieved by designing these
structures to a larger earthquake return period and consequently to higher values of seismic input. This
approach corresponds to designing them to a lower probability of damage and collapse in the event of
future earthquakes.

Similarly, different levels of probability of occurrence of earthquakes can also be used for different design
philosophies. For regular structures, the choice of an earthquake level with a very low probability of being
exceeded is usually associated with a design aimed at avoiding structural collapse, and thus human
casualties, even if the structure undergoes major damage and has to be rebuilt. For earthquake levels with
higher probability of occurrence, and that may thus occur more often during the lifetime of the structure,
the design goal is not to avoid collapse but rather to guarantee that no substantial damage occurs and that
the structure maintains its serviceability.

Usually, the response spectra are presented in normalized form, as is the case of the normalized elastic
response spectrum of Eurocode 8. It is normalized to the peak ground acceleration (ag), i.e. it is
independent of ag and so can be used for different values of the maximum expected acceleration for the
site. This approach allows for the use of the same spectra for different conditions of severity of the ground
motion. In other words, it enables the consideration of earthquakes corresponding to different return
periods and thus to different acceptance of seismic risk.

According to Eurocode 8 and other national regulations, the elastic response spectrum to be used for
design purposes depends on several parameters such as the seismic zone, the type of seismic action, the
local soil conditions and the viscous damping ratio of the structure.
The seismic zone can be characterized by means of the severity of the seismic action. This
characterization is accomplished by normalizing the response spectra to a certain level of ag. Usually, the
response spectrum for the vertical motion is defined as a percentage of the response spectrum for the two
orthogonal horizontal directions. In Eurocode 8 the suggested percentage is 70%.

The maximum acceleration to be used in each region in Europe is defined according to microzonation
studies for each zone, depending on the local seismic hazard parameters. It is the responsibility of the
National Authorities.

The normalized elastic response spectrum e (T) (Figure 8) is defined by means of four parameters, o,
T1 T2 and k, according to the following expressions:

0 < T < T1 e (T) = 1 + T/T1 (o - 1)

T1 < T < T2 e (T) = o

T2 < T e (T) = (T2/T)k o

where

T is the natural vibration period of the structure, or the inverse of the natural frequency (Hz)

o is the maximum value of the normalised spectral value assumed constant for periods between T 1 and T2

k is an exponent which influences the shape of the response spectrum for vibration periods larger than T 2.

The values of the transition periods T1 and T2, also known as the inverses of the corner frequencies,
depend essentially on the magnitude of the earthquake and on the ratios between the maximum ground
acceleration, ground velocity and ground displacement.
The basic values presented in Eurocode 8 [1] apply to the ground motion at bedrock or in firm soil
conditions. If the soil characteristics differ from the ones considered, other values for the parameters can
be chosen in such a way that the shape of the response spectrum is modified accordingly. Eurocode 8
considers three different soil profiles (A, B and C). For each soil profile different parameters (o,
T1 T2 and k) apply. The local response spectrum, s (T), can be obtained, correcting the elastic response
spectrum by a soil parameter S, which is also dependent on the soil profile.

s (T) = S e (T)

Although the basic form of the response spectrum is uniform, and is common to the designers in every
European Community country, the parameters that define the response spectrum are also the
responsibility of each National Authority. The parameters can vary from region to region even in a single
country. This variation is due to the fact that each European region has different seismicity.

The o value is the maximum spectral amplification. It is dependent on the selected probability of being
exceeded for the considered peak ground acceleration, on the damping ratio, on the duration of the ground
motion, and on its frequency content. According to Eurocode 8, for a 20 to 30 second earthquake and 5%
damping, the value of o = 2,5 corresponds to a probability of not being exceeded of between 70 and 80%
[1].

The exponent k is dependent on the frequency content and the selected probability of being exceeded. It
describes the shape of the response spectrum for the higher periods (lower frequencies).
The use of the elastic response spectrum, simultaneously with linear elastic design, does not take into
account the capability of a structure to resist seismic actions beyond the elastic limit. If it can be assumed
that the structure will behave linearly for small earthquakes, for larger earthquakes it would be almost
impossible and non-economical to design structures based on the assumption of linear behaviour. For
larger earthquakes it should be assumed that the structure has a certain capacity to dissipate the energy
input by the earthquake by means of non-linear behaviour, even if that implies the existence of structural
damage although guaranteeing that collapse is avoided.

Thus, for design purposes, and to avoid the necessity of performing non-linear analysis, the concept of
structural behaviour factor (q) is introduced, to correct the results obtained by linear analysis and obtain
an estimate of the non-linear response. These behaviour factors, which will be presented in more detail in
other lectures, take into account the energy dissipation capacity through ductile behaviour. Thus they are
dependent on the materials, type and characteristics of the structural system and the assumed ductility
levels. Eurocode 8 defines the q values to be adopted in the case of steel structures according to criteria
that will be presented in later lectures.

Based on the q factors, it is possible to define the linear analysis design response spectra that can be used
for design purposes by means of linear analysis.

The linear analysis design response spectra is defined in Eurocode 8 as follows:

0 < T < T1 (T) =  S [1 + T/T1 ( o/q - 1)]

T1 < T < T2 (T) =   S  o/q

T2 < T (T) = (T2/T)k   S o/q

where

T, o, T1, T2 and k have the same meaning as above.

 is the ratio of the peak ground acceleration to the acceleration of gravity.

 is a conservative factor for damping ratios different from 5%.

q is the behaviour factor which can depend on T.

The influence of the structural damping ratio is obtained by means of:

 =  (5 / );  > 0,70

where  is the value of viscous damping ratio as a percentage.

According to Eurocode 8, if there is a possibility of two earthquake sources affecting a given site, the use
of two different response spectra may be necessary to quantify the seismic input and response [1]. This
possibility may arise for sites that may be affected by very large magnitude earthquakes with large
epicentral distances and simultaneously by smaller but nearby earthquakes. In that case, although the
ag or  o values may be quite similar, the shapes of the two corresponding spectra may vary substantially
(Figure 9). As a result, some structures may be more affected by one of the earthquakes, whereas other
structures may be more affected by the other one.

If a more sophisticated approach is required, and non-linear analysis is to be performed, or if alternative


design is to be made, the use of earthquake time-histories, or records of ground acceleration, is necessary.
When insufficient previously recorded earthquake accelerograms are available or when they do not
belong to the same seismic environment, artificially generated earthquakes may be used. There are several
alternative methodologies for generating artificial earthquakes. The only constraint is that the generated
histories shall be consistent with the response spectrum corresponding to the case under study. The same
applies to the use of power spectra to represent the seismic action.

As a final observation on the characterization of the seismic motion, the effects of the spatial variability of
the seismic motion should be considered. The seismic input may be different from support to support. The
differences are due to several factors such as the overall dimensions of the structure, the large distances
between two supports of the same structure, or the fact that a structure may have different foundation
conditions, both in terms of soil or foundation types. In this case a spatial model of the seismic action has
to be used, taking into account a model of the wave propagation.

4. FINAL REMARKS

The social consequences of earthquakes, in terms of human casualties and injuries and direct and indirect
economic losses justify the need to be prepared for earthquakes. Earthquakes are still difficult to predict
and, even if they could become predictable, would pose a threat to buildings and other structures. Thus,
being prepared for earthquakes consists mainly in proper structural design procedures for seismic loading.
To achieve a correct design procedure and thus diminish the seismic risk, it is necessary in the first place,
to have a correct knowledge of the seismic input, or the seismic hazard. Simultaneously with the study of
the behaviour of structures when subject to seismic loading it is thus fundamental to study the seismic
motion, its origin, and the parameters that most influence the characteristics of the motion.

5. CONCLUDING SUMMARY

 Earthquakes are natural phenomena that have caused tremendous losses of lives and goods
worldwide including in some large areas of Europe.
 To design structures that can resist earthquakes requires an understanding of the seismic
hazard.
 "Measuring" the earthquake can be achieved by means of different parameters such as
magnitude, intensity, peak ground acceleration, power spectrum and response spectrum.
Parameters that influence the characteristics of the earthquake motion and its response
spectrum are the duration and frequency content of the motion and the local soil conditions.
 The response spectrum approach presented in Eurocode 8 which can be used for structural
design takes into account a probabilistic approach of the definition of the seismic motion [1].

6. REFERENCES

[1] Eurocode 8: "Structures in Seismic Regions - design", Commission of the European Communities,
Report EUR 12266, 1989.

7. ADDITIONAL READING

1. Clough, R. W. and Penzien, J., Dynamics of Structures,

McGraw-Hill - International Student Edition, 1975.

2. Gere, J. M. and Shah, H. C., Terra Non Firma - Understanding and preparing for earthquakes,
Stanford Alumni Association, Stanford, USA, 1984.
3. Catalogue of European earthquakes with intensities higher than 4,

Commission of the European Communities, Report EUR 13406, 1991.

4. Dowrick, D. J., Earthquake Resistant Design, Wiley and Sons, 1987.

Previous | Next | Contents

Previous | Next | Contents

ESDEP WG 17

SEISMIC DESIGN

Lecture 17.3: The Cyclic Behaviour of

Steel Elements and Connections


OBJECTIVE/SCOPE

To give basic knowledge about the ductility resources of steel members and connections under cyclic
loading.

PREREQUISITES

None.

RELATED LECTURES

None.

SUMMARY

After a brief introduction and a description of the cyclic behaviour of the material this lecture examines
the ECCS recommended testing procedure for assessing the behaviour of structural steel elements under
cyclic loads in the context of earthquake resistant design. A description of the loading history and the
interpretative parameters of the recommended testing procedure and their use is also presented.

An overview of recent European research work is given. Experimental results on the cyclic behaviour of
bracing elements, beams, columns, beam-to-column connections and shear links for eccentrically braced
steel frames are presented to illustrate typical behaviour and physical phenomena related to failure modes
and deterioration of resistance. A comparison between the cyclic behaviour of different detailed
connections (fully welded joints, bolted joints using angles, cover plates, flanges) is discussed.

1. INTRODUCTION

Today steel rolled products, such as H or tube sections, are available in a wide variety of types and
dimensions, larger than in the past. They may be used to produce a wide range of structural elements and
connections.

Steel elements have the advantage that huge elements can easily be constructed. They may be considered
the most appropriate building material to ensure the seismic resistance of large structures. The
performance and ductility of the structural elements and connections may be affected by many factors.

Structural steel elements often have high slenderness and slender cross-section due to high strength, and
various types of buckling may occur, such as flexural buckling of the whole element, lateral-torsional
buckling and local buckling of plate elements constituting the element.

In the Eurocodes, cross-sections are classified with respect to the proportions and loading conditions of
each of their compression elements. Compression elements include every element of the cross-section
which is either totally or partially in compression. Under cyclic loading the increase of the width-to-
thickness ratio of the compression elements lowers the resistance, the ductility and the dissipated energy
arising in the various types of buckling.

The occurrence of buckling brings a sudden reduction of the load-bearing resistance of the member. Even
when buckling does not cause the immediate failure of the element, careful consideration should be given
in design to the prevention of buckling because the response of the structure to an earthquake often
becomes unstable due to the buckling of some elements.

In addition to analysis of the individual elements, the stability of the frames as a whole should also be
analysed. In general second order effects included in the global analysis of sway frames produce a
progressive decrease of resistance and a reduction of the dissipated energy and ductility. For that reason,
the second order effects have always led to an overdimensioning of columns for frame-type buildings
erected in seismic zones.

Seismic actions produce deformations with relatively few repetitions of the action. Deformations of fairly
large amplitude occur at fairly low speeds. These deformations exhibit cyclic characteristics which may
produce low cycle fatigue phenomena of structural elements and connections but rarely their failure.
However, the possibility of damage from element failure should be considered in design against external
cyclic loadings such as those produced by earthquakes.

Over the last twenty years tests have been performed in universities and research centres to obtain a better
understanding of the seismic behaviour of structural steel elements and to characterise their ability to
deform in the inelastic range. However, the testing procedures and the interpretative parameters of the
tests have differed from one researcher to another, making qualitative and quantitative comparison
difficult in some cases.

The European Convention for Constructional Steelwork (ECCS), through its Technical Committee 1 -
Structural Safety and Loadings [1], has suggested a testing procedure for assessing the behaviour of
structural steel subassemblages under cyclic loads [1]. This procedure is intended to be a defined method
of experimental testing for structural elements or complete structures such that comparisons of results
obtained by different authors are possible. The procedure is also intended to enable the assessment of the
seismic behaviour of steel elements based on cyclic quasi-static tests using a specified loading history.

2. DUCTILITY

In seismic design it is very important to assess the ability of a structure to develop and maintain its
bearing resistance in the inelastic range. A measure of this ability is ductility, which may be referred to
the material itself, to a structural element, or to a whole structure. These three kinds of ductility are very
different in their numerical values, and each one plays a significant role in seismic design.

Material ductility - e, measures the ability of the material to undergo large plastic deformations. A high
value of e characterises a ductile material, a low value a brittle one.

Structural element or joint ductility -  , characterises the behaviour of a member or joint, and
particularly its ability to transmit stresses in the elastoplastic range without loss of resistance. For
instances a frame structure cannot show a ductile behaviour if the plastic hinges are not able to
redistribute the bending components.

Structural ductility -  , is an index of the global behaviour of the structure, i.e. the ability of a structure
to deform in the inelastic range after some of its parts have exceeded their linear elastic range.

The ductilities e,  and  must meet the condition:

e >  > 
3. MATERIAL

The steel used in an earthquake resistant structure must, of course, be of good quality. In addition to the
general requirements for the material, the steel must have adequate ductility.

Figure 1 shows the stress-strain relationship of a structural steel under uniaxial hysteretic loops. In the
first loading, the upper yield point, the lower yield point, the plateau and strain hardening appear clearly.
In subsequent loadings, these properties disappear and the proportional limit markedly decreases due to
the well-known Bauschinger effect.

Steels are usually considered to possess such prominent properties as yield point, plateau and strain
hardening. For repeated loading beyond the elastic range, however, the stress-strain diagram with no
plateau should be applied instead. Attention should be given to the possibility that the reduction of the
proportional limit in each loading cycle may cause direct structural failure (for instance, buckling),
increase of deformations and reduction of rigidity of structural elements, connections and cross-sections
of elements.

Steel is an alloy of iron with carbon and various other elements. The carbon exerts the most significant
effects on the microstructure of the material and its properties. Changes in the carbon percentage affect
the strength, toughness and ductility of the steel.
Steels normally used in structures are excellent materials possessing a high ductility in the direction of
rolling. Low steel grades of steel show better ductility than high grades.

The ductility of steel may be described as its post-elastic behaviour and may be measured, provided that
the stress-strain relationship is known, as the ratio:

ductility = (deformation at failure) / (deformation at yield)

The numerical value of ductility is usually represented by the ductility factor, e, and depends on the
origin considered for the deformation at failure. In general it can be defined by the ratio of the maximum
deformation at the beginning of the cycle, u, to the yield deformation, y:

e = u / y

For cyclic loading and for a specified loading history, u may be defined as the maximum deformation
from the initially undeformed material u , or the deformation from the beginning of a cycle to the new
maximum u" (Figure 1). The last definition appears more meaningful for the assessment of cyclic
behaviour.

The ductility in the material is desirable and necessary since the ductility of structural elements and of
whole structures depends on the material characteristics. However, the possibility of brittle behaviour
must be carefully guarded against by proper detailing and good workmanship.

The area within the hysteresis loop corresponds to the dissipated energy of the loop.

4. LOADING HISTORIES

Various types of loading histories have been idealised in cyclic tests to evaluate the resistance-
deformation characteristics of structural subassemblages. The most commonly used are (Figure 2):
(a) no force reversal.

(b) force (F) reversal, but no deformation () reversal.

(c) partial deformation reversals.

(d) full deformation reversals.

(e) random.

The type of load reversal has an important influence on the cyclic behaviour. Full deformation reversals
cause more deterioration of the resistance of the element than partial deformation reversals, see Figure 3.
Cyclic loadings produce much more deterioration of the resistance than monotonic loadings.
The selection of the loading history to use in the assessment of the seismic resistance of steel
subassemblages is a very difficult task because it is impossible to know the real loading history in future
earthquakes. However certain factors can be taken into account in choosing a loading history.

In general, displacement increase should be preferred to load increase because the resistance of a
structural element may decrease after a few cycles due to buckling phenomena for instance. In this case, if
load increase is applied the test cannot be controlled and it is probably best to discontinue it. However, it
may be of interest to proceed with the test considering that the element is only a part of the structure and
its decrease in resistance does not necessarily mean an important decrease in resistance for the structure.

As indicated above, full deformation reversal causes more deterioration of resistance than partial
deformation reversals. Full deformation reversal is probably the most generally used loading for assessing
the resistance to damage of the anti-seismic parts of a structure. The type of reversal used in tests however
should be defined considering that the structural element is part of the whole structure and should be
designed to resist both static and seismic actions.

The number of cycles at a constant maximum displacement should also be defined. The number of
repetitions defined at the same displacement should not be too high in order to avoid low cycle fatigue
phenomena, since the number of high peaks of displacement caused by real earthquakes is generally low.

5. ECCS TESTING PROCEDURE

The procedure for assessing the behaviour of structural steel elements under cyclic loads recommended
by the ECCS [1] can be applied to plane or three dimensional tests and may include preliminary
monotonic displacement tests. This procedure is designated the complete testing procedure. If monotonic
tests are omitted it is designated the short testing procedure.

5.1 Complete Testing Procedure

This procedure includes three tests performed on different specimens. The first and second tests impose
displacement increasing monotonically in the tension and in the compression range respectively. The
positive and the negative reference elastic load Fyand the corresponding reference elastic
displacement y are obtained from the recorded force-displacement curve. The reference elastic load is
defined as the intersection between the tangent modulus Et at the origin of the force-displacement curve
and the tangent that has a slope of Et/10 as indicated on Figure 4c.
Other conventional definitions of Fy may be used, such as (a) the value corresponding to the 0,2% offset
load at some point in the tested specimen (Figure 4a), or (b) the maximum load (Figure 4b). Definition (a)
ignores the post-elastic reserves of the specimen and definition (b), in spite of its interest in the buckling
context, may correspond to exaggerated deformation of the flexural behaviour of beams or joints.

The definition of Fy recommended by the ECCS (Figure 4c) covers many cases and types of behaviour
and avoids some disadvantages of the definitions (a) and (b).

The third test is a cyclic test with increasing displacement as follows:

 one cycle in the interval [y+/4; y-/4],


 one cycle in the interval [2y+/4; 2y-/4],
 one cycle in the interval [3y+/4; 3y-/4],
 one cycle in the interval [y+; y-],
 three cycles in the interval [(2+2n)y+; (2+2n)y-] with n = 0,1,2,3...

The end of the test is not defined beforehand. For research purposes the test will probably be continued as
far as possible in order to obtain the maximum information. On the other hand a design engineer will
probably stop the test as soon as the code requirements are reached.

5.2 Interpretation of Tests

Several problems are raised when it is necessary to compare different test results due to the diversity of
the interpretative parameters used.
The recommendations of the ECCS use a standardisation of the interpretative parameters which are
established in ratios which are meaningful to the engineer [1]. The suggested parameters are normalised
with reference to those corresponding to a linear elastic-ideal plastic behaviour.

Since the behaviour of the element in the tension and in the compression range may be different, the
parameters are evaluated in these two ranges. The quantities used in the ratios are deduced from the force-
displacement curve and are obtained for cycles with displacements larger than the reference elastic
displacement. The proposed parameters for a typical cycle (Figure 5) are:

 Full ductility ratio:

i+= i+/ [i+ + i- - y-] i-= i-/ [i- + i+ - y+]

 Resistance ratio:

i+= Fi+/ Fy+ i-= Fi-/ Fy-

 Rigidity ratio:

i+= tg i+/ tg y+ i-= tg i-/ tg y-

 Absorbed energy ratio:


i+= Ai+/ [i+ + i- - y+ - y-]Fy+

where

i+ (i-) is the value of the maximum positive (negative) displacement in the ith cycle.

y+ (y-) is the value of the positive (negative) reference elastic displacement.

i+ (i-) is the value of the maximum displacement in the positive (negative) force range in the ith cycle.

Fi+ (Fi-) is the value of the positive (negative) force corresponding to the i+ (i-) in the ith cycle.

Fy+ (Fy-) is the value of the positive (negative) reference elastic force.

tg+i(tg-i) is the value of the slope of the tangent to the force-displacement curve when F changes from
negative (positive) to positive (negative) at the ith cycle.

tg+y(tg-y) is the value of the slope of the tangent at the origin of the force-displacement curve for
increase in positive (negative) direction.

Ai+ (Ai-) is the area under the positive (negative) force range of the half cycle in the force-displacement
curve.

In general the behaviour of the element is better when its behaviour is near to the reference linear elastic-
ideal plastic behaviour (values of parameters near to one).

Small values of the parameters, i.e. much less than 1, may be assumed as an index of the end of the test
because they indicate large losses of ductility, resistance, rigidity and absorbed energy.

The parameters proposed by the ECCS have the advantage of assisting the quantitative analysis of the
cyclic behaviour of structural elements [1]. They can also be considered as practical parameters for the
definition of code acceptance criteria.

6. BRACING ELEMENTS

A typical plastic hysteretic behaviour of a bracing element under repeated loading is shown in Figure 6.
Slide 1 shows its mode of failure. The force-displacement curve was obtained from experiment, in which
an element made up of back-to-back angles 80 x 80 x 8 and slenderness ratio of 145 was subjected
through pinned ends to repeated tension and compression. The strain amplitude was gradually increased
in each loading cycle approximately following the ECCS short testing procedure [1].
Slide 1 : Mode of failure of a bracing element made up of back-to-back angles 80 x 80 x 8 and
slenderness ratio of 145.

The initial buckling load corresponds to point A. Beyond point A on the force-displacement curve, the
bracing element suffers a considerably loss of resistance as it buckles. This phase of the hysteretic
behaviour A-B is dominated by the plastic interaction between the bending of the brace element due to
the P- effect induced by the compressive force. It is characterised by large lateral deflections and
decreasing load. If the increment in the displacement is reversed, the force-displacement curve
corresponds to the elastic recovery B-C with a brace lengthening C-E. In the case of bolted connections, a
slip occurs in the force-displacement curve, zone C-D. In zone E-F a plastic interaction of axial force and
bending moment occur with a decrease of the lateral deflection. At point F the element is fully
straightened. The zone F-G is characterised by the plastic elongation of the element. Reversing the
direction of the displacement results in the elastic unloading of the brace, zone G-H.

Figure 6 also shows a decrease of the ultimate compressive load with the application of the cyclic
displacement, as a consequence of the Bauschinger effect. This decrease can also be a consequence of the
brace curvature. In general, after an initial buckling cycle, the brace develops a residual curvature which
is not completely removed by the subsequent tensile yielding. The brace behaves as an element with an
initial curvature.

The hysteretic behaviour of a bracing element is affected by its slenderness ratio. Braces with small
slenderness ratio absorb more energy than more slender ones, as can be seen by comparing their
hysteresis loops. In general, bracing elements with large slenderness ratios show a more rapid
deterioration in their compressive bearing resistance than those with smaller ratios.
For bracing elements for which the strength in compression is accounted for in the evaluation of the
lateral stability of the frame (K bracing for instance), it is advisable to limit the reference slenderness ( )
of the brace to values in the range 1,0 - 1,5. is defined in Eurocode 3 [2] as:

where Nc is the compression resistance and Ncr is the Euler critical load. For the above values of ,
mean values for slenderness  are equal to 94-140 for steels Fe E 235 and 76-114 for steels Fe E 355. In
general no requirements are necessary for X bracing or truss bracing because only the tension diagonal is
accounted for in the evaluation of the seismic resistance.

According to Eurocode 8 [3] the connections of bracing elements to other elements shall fulfil the
overstrength condition:

Rd  1,20 Npd

where Rd is the resistance of the connection and Npd is the ultimate resistance of the connected part
according to Eurocode 3 [2]. This condition ensures that the connected element fails before the
connection.

The hysteretic behaviour of the brace is affected to some extent by the shape of the cross-section. Figure 7
shows the hysteresis loops of a channel bracing element under repeated loading. Slide 2 shows its mode
of failure. The cross-sectional shape affects the susceptibility of the brace to lateral-torsional buckling and
local buckling and, as a result, the compressive carrying resistance.
Slide 2 : Mode of failure of a channel bracing element under repeated loading. The cross-section shape
affects the susceptibility of the brace to lateral-torsional buckling and local buckling and, as a result, the
compressive resistance.

In general, rolled steel sections as currently produced exhibit local buckling at extremely large lateral
displacements. Experimental studies performed by several authors indicate progressively poorer
performance of cross-sectional shapes in the sequence: tubes, wide flanges, tees, double channels and
double angles. The poor performance of tees and double angles in comparison with tubes and wide
flanges can be assigned to their geometric proportions and single symmetry. Tees and double angles
buckle in the direction perpendicular to their axis of symmetry causing flexural and lateral-torsional
buckling. Consequently they have a lower carrying resistance than that which would develop in pure
flexural buckling.

Built-up braces should be designed as a single element. It is important to observe the structural rules for
detailing the design in order to avoid early buckling of individual elements under low load.

The reinforcement of support points must not be forgotten in order that bracing elements may fulfil their
expected purpose. If the ends of an element can displace easily, the stability of the whole building must be
considered. Generally, bracing elements are connected by a gusset plate which has a low flexural rigidity.
For this reason, the gusset plate may require reinforcement to increase it's bending resistance.

7. BEAMS AND COLUMNS

Figure 8 shows a force-displacement hysteresis curve obtained from an experiment on an I cantilever


beam subjected to repeated loading according to the ECCS short testing procedure [1].
Experimental investigations carried out on cantilever beams under repeated and reversing loading have
shown that the development of local buckling in the flanges does not signal an immediate loss of the
moment resistance. The beams are able to sustain loads substantially higher than those that cause initial
flange buckling. This behaviour is attributed to the considerable post-buckling strength of the plate
elements. However, after the occurrence of the maximum load in the subsequent load cycles, the moment
resistance deteriorates. This deterioration is higher with increasing width-thickness ratio (b/t) of the
flanges as a consequence of the early occurrence of local instability in the flange elements.

The severe distortions of the flanges tend to induce torsional displacement of the section (Slides 3 and 4).
They are associated with a lower load that would develop in pure flexural buckling. This effect is likely to
contribute to the somewhat poorer performance of H and I beams as compared to box shaped cross-
sections as shown in Figure 9 and Slides 5 and 6. For this reason, unless supports in the lateral direction
are provided, the use of a box section is preferable. Similarly the behaviour of truss beams is improved by
using steel tube with high torsional rigidity as flange elements.
Slide 3 : The severe distortions of the flanges of H and I sections tend to induce torsional displacement of
the section.
Slide 4 : The severe distortions of the flanges of H and I sections tend to induce torsional displacement of
the section.
Slide 5 : Box sections are torsionally stiff and are therefore much less susceptible to lateral-torsional
buckling.
Slide 6 : Box sections are torsionally stiff and are therefore much less susceptible to lateral-torsional
buckling.

To allow the development of a plastic hinge with a very good rotational capacity in rolled beams with H
and I cross-section, the following condition of the flange is required (b flange width, t flange thickness):

This condition is generally satisfied by rolled steel sections currently available. This limiting ratio ensures
that the flanges can be compressed uniformly without buckling up to strains in the strain hardening range
of the material. The increase of stability is in general accompanied by an increase of the ductility of the
beam.

Limited information is available on the cyclic behaviour of beams with b/t ratios exceeding the limiting
value. However, the cyclic behaviour and strength of these beams is similar to those with b/t ratio of
flanges less than this limit. However flange buckling has been observed at a moment slightly higher than
the plastic moment.

Further research is necessary in order to verify the limiting width-thickness ratio for plates under cyclic
bending.
For beams with cross-sections having different ultimate characteristics in the two directions, the rotational
capacity and the ductility factor may also be different in the two directions. T-sections, for example, have
different rotational capacities in the two principal directions of the bending.

In frames, in order to guarantee sufficient hysteresis rotation capacity of beams under the full plastic
moment action effects, the following inequalities shall be verified at the locations where the formation of
hinges is expected according to Eurocode 8 [3]:

 1,0

where

N, M are the action effects taking account of the behaviour factor q.

Npd, Mpd are the ultimate resistances according to Eurocode 3.

Vo is the shear force due vertical loads.

VM is the shear force due to the resisting moments of the beam and its extremities.

Vpd is the shear resistance of the beam according to Eurocode 3 [2].

Some experimental information on the behaviour of columns under repeated bending with a constant axial
force is available. It shows that, where there is a large axial force, the height of the first loading curve is
low and the gradients of the curves are negative after the attainment of the maximum load in each loading
cycle (Figure 10). This effect is commonly referred to as the P- effect. It should be noted, however, that
the load-carrying resistance increases in each loading cycle because of the accumulated compression
strain hardening under repeated bending and constant force. The strain accumulation caused by cyclic
bending would reduce the rotation capacity of the section. The extent of the reduction has not been
thoroughly investigated. It is not yet known the extent to which this reduction affects the load-carrying
resistance of columns.
The existence of axial force in the columns leads to a more rapid decay of the load-carrying resistance
owing to more extensive buckling in comparison to beams.

8. CONNECTIONS

There are many types and varieties of connections, and each has different rotational characteristics that
affect the frame behaviour. Butt welding, fillet welding, bolting, and riveting may be employed for
aseismic connections, either individually or in combination. As fully bolted or riveted connections tend to
be large and expensive, fully welded connections or a combination of welding and bolting are the most
frequently used. Bolts have the advantage of providing more damping to frames than welds.

Connections should be designed to make fabrication and erection of the framework as simple and rapid as
possible.
Conclusive design criteria for beam-to-column joints are not yet available for seismic conditions. Until
the recent past relatively few cyclic load tests had been performed on joints commonly used in Europe. At
present many experimental investigations are in progress in different European laboratories. They deal
with cyclic behaviour of rigid and semi-rigid joints, both for bare steel and composite constructions.

Preliminary research to investigate the influence of detailing of the joint was performed by Ballio,
Mazzolani et al on fourteen specimens [4, 5]. The connection types were in compliance with the
technology commonly used in Europe for rigid and semi-rigid joints. The experiments followed the ECCS
recommended testing procedure for short tests [1]. The specimens were grouped into four main categories
(Figure 11):
Type A - This type of connection is made using three plate splices which are welded to the column and
bolted to the flanges and to the web of the beam. The basic type A1 is varied by the introduction of
diagonal stiffeners in the column web (A2, A4) or reinforcing plates in the beam flanges (A3, A4).
Type B - Angle splices are bolted both to the column and to the beam. The basic type B1 is varied by
stiffening the column (B2, B4) or the angle splices connected to the beam flanges (B3 - B4).

Type C - End plate joints with rigid column stub. Variations from the basic type C1 are derived by
introducing stiffeners in the beam web (C2, C3, C4) or increasing the thickness of the end plate (C3, C4).

Type D - Fully welded connections of basic type (D1) or varied with reinforcing plates on the column
web (D2).

Comparison between the results (Figure 12) indicates the role played by the detailing of the connections
under alternating loading conditions.
Comparing for instance A4 to A3, the introduction of a diagonal plate to stiffen the central panel of the
column (Slides 7 and 8) reduced the energy dissipated, increased the strength, and the collapse became
brittle because the failure occurred at the maximum load. The opposite behaviour was found for A1 and
A3 which collapsed in a ductile manner.

Slide 7 : The introduction of a diagonal plate to stiffen the central panel of the column reduces the energy
dissipated, increases the strength, and results in brittle collapse because the failure occurs at maximum
load.
Slide 8 : The introduction of a diagonal plate to stiffen the central panel of the column reduces the energy
dissipated, increases the strength, and results in brittle collapse because the failure occurs at maximum
load.

The stiffening elements placed under the column flange to control the deformation produced by the
tension force of the angle profile bolts (Slides 9 and 10) increased the resistance of connection B4
compared to B3, for instance. The introduction of a triangular plate in the angle connecting the beam and
column flange also produced an increase of energy and resistance.

Slide 9 : Stiffening elements placed under the column flange to control the deformation produced by the
tension force of the angle profile bolts increases the resistance of the connection.
Slide 10 : Stiffening elements placed under the column flange to control the deformation produced by the
tension force of the angle profile bolts increases the resistance of the connection.

The addition of the beam web-flange stiffeners (compare C2 to C1 in Slides 11 and 12) reduced the
energy and increased the resistance. The increase of thickness of the end-plate in C3 and C4 or the
introduction of partial or full stiffening plates in the beam improved the load level, but not sufficiently to
compensate for the energy dissipated by C1.
Slide 11 : The addition of beam web-flange stiffeners reduces the energy and increases the resistance.

Slide 12 : The addition of beam web-flange stiffeners reduces the energy and increases the resistance.

The stiffening of the column panel in D2 produced a decrease of energy absorption and an increase of
load level reached in comparison to D1 (Slides 13 and 14).

Slide 13 : The stiffening of the column panel produces a decrease of energy absorption and an increase of
load level.
Slide 14 : The stiffening of the column panel produces a decrease of energy absorption and an increase of
load level.

Based on these tests, some general qualitative indications regarding detailing may be drawn:

 If stiffeners are added to the parts of the connection which are most responsible for its
flexibility, the amount of energy absorption decreases but the load level is increased.
 If elements are added to a joint which do not substantially modify the deformation mechanism
but increase the local strength of the structural elements then there will be an increase of
energy absorption and load level provided that the collapse is ductile.

For this type of connection the plastic rotation of the beam is mainly developed by the extension of plastic
deformation near the connection. Generally, in order to control the extension of the plastic region in the
element in the vicinity of the connection, the beam-to-column connection must have an ultimate bending
moment which is greater than the full bending resistance of the attached element. For this reason
Eurocode 8 [3] requires that the resistance of the connection be greater than the resistance of the adjacent
connected element:

Rd  1,20 Rfy

where Rd is the resistance of the connection according to Eurocode 3 [2] and Rfy is the yielding resistance
of the connected part. Connections made by means of butt-welds or full penetration groove welds are
deemed to satisfy this overstrength criterion.

Beam-to-column connections are one of the most common types of connections in steelwork. However
other types of connections may occur in steel frames. In eccentrically braced frames (Figure 13) the axial
forces in the brace are transferred to another brace, or to a column through shear and bending in a short
segment of beam, usually called the active link. Its behaviour is strongly dependent on its length. If it is
sufficiently long, plastic moment hinges form at both ends of the link. On the other hand, if this link is
short it tends to yield in shear with smaller end moments.
Active links equal to or shorter than b* (Figure 13) will yield predominantly in shear, and are called shear
links. Links that are somewhat longer experience substantial moment-shear interaction. The end moments
of the long links will approach the plastic moment resistance of the beam, and moment hinges will form
at the ends of the links. Such links are referred to as moment links.

For moment links a large increase in shear can take place with only a small change in moment.
Conversely, for shear links the shear resistance remains essentially constant for a considerable range of
end moment.

Based on the results of the investigation performed by Popov et al on the seismic behaviour of active
links [6, 7, 8], some general conclusions can be drawn from the hysteretic behaviour of this type of
connection:

 Inelastic shear is more efficient than inelastic web buckling for energy dissipation.
 Stiffening improves the energy dissipation capability of an active link by delaying the onset of
inelastic web buckling (Figure 14). Stiffening slows the degradation of load-carrying resistance in
a link by controlling the amplitude of out-of-plane displacement of the web.
 Interaction of web and flange buckling fields causes a more severe degradation of resistance
than either of the modes acting alone.
According to Eurocode 8 [3], in order to guarantee the formation of a shear mechanism in the active link
with full deformation capacity, the full resistances to the action-effects other than shear are restricted to
values as follows:

 1,00  0,70  0,15


where V, M and N are the action-effects and the index pd denotes the respective ultimate resistance.

9. CONCLUDING SUMMARY

 In seismic design it is very important to assess the ductility of the structure.


 Many factors may affect the performance and ductility of the structural elements and
connections, such as, slenderness, cross-section shape, second order effects and detailing.
 Low carbon steels normally used in structural work are excellent materials possessing a high
ductility in the direction of rolling.
 The hysteretic behaviour of bracing elements is affected by their slenderness ratio. Increase of
slenderness produces decrease of the absorbed energy.
 The ductility of beams and columns is greatly affected by the width-to-thickness ratio of the
compression elements and by the level of axial load.
 The detailing of connections may have a substantial effect on their flexibility, energy absorption,
strength and ductility.

10. REFERENCES

[1] Recommended Testing Procedure for Assessing the Behaviour of Structural Steel Elements under
Cyclic Loads ECCS Publication No45, 1986.

[2] Eurocode 3: "Design of Steel Structures": ENV 1993-1-1: Part 1.1, General rules and rules for
buildings, CEN, 1992.

[3] Eurocode 8: "Structures in Seismic Regions - Design", CEN (in preparation).

[4] Ballio, G., Calado, L., De Martino, A., Faella, C., Mazzolani, F. M. (1987) Cyclic Behaviour of Steel
Beam to Column Joints: Experimental Research, Costruzioni Metalliche No. 2. pp 69-90.

[5] Popov, E. (1980) Seismic Behaviour of Structural Subassemblages. ASCE Journal of the Structural
Division, ST7, Page 1451-1470.

[6] Popov, E. (1980) An Update on Eccentric Seismic Bracing. AISC Engineering Journal n. 3, Page 70-
71.

[7] Popov, E. and Roeder, C. (1978) Design of Eccentrically Braced Steel Frame. AISC Engineering
Journal, n.3, Page 77-81.

11. ADDITIONAL READING

1. 'Earthquake Resistant Design for Engineers and Architects', David J. Dowrick, John Wiley & Sons,
1987.
2. 'Design of Earthquake - Resistant Buildings', Minoru Wakabayashi, McGraw-Hill, Paris, 1986.
3. 'Study on Design of Steel Building in Earthquake Zones' ECCS Publication no. 47, 1986.

Previous | Next | Contents

Previous | Next | Contents


ESDEP WG 17

SEISMIC DESIGN

Lecture 17.4: Structural Analysis for

Seismic Actions

OBJECTIVE/SCOPE

To give an overview of the methods used for the analysis of structures under seismic actions.

PREREQUISITES

Basic knowledge of structural analysis and structural dynamics

RELATED LECTURES

Lecture 17.2: Introduction to Seismic Design - Seismic Hazard and Seismic Risk

Lecture 17.3: The Cyclic Behaviour of Steel Elements and Connections

SUMMARY

The lecture briefly presents the methods stipulated by modern design codes for the analysis of structures
under seismic actions. Time-domain methods are briefly described and the scope of their application is
specified. Emphasis is given to the response spectrum method as the standard procedure proposed by, for
example, Eurocode 8 [1]. In addition, a simplified response spectrum method for regular buildings is
presented. Finally inelastic behaviour and its role in design under seismic actions is discussed.

1. INTRODUCTION

Several methods are available for the structural analysis of buildings and other civil engineering works
under seismic actions. The differences between the methods lie (a) in the way they incorporate the
seismic input and (b) in the idealization of the structure. All methods of analysis must serve the current
design philosophy for seismic actions which requires that a structure must not collapse and must retain its
structural integrity under the so-called "strong" earthquake. The structure also must be protected against
damage and limitations of use under the so-called "moderate" earthquake. To avoid collapse, the structure
is allowed to develop plastic zones in which seismic energy is dissipated.

Details of the basic requirements of seismic behaviour of structures, and the criteria needed for ensuring
compliance with these requirements, can be found in all modern seismic design codes, e.g. Eurocode 8
[1].

2. DIRECT METHODS OF DYNAMIC ANALYSIS (TIME INTEGRATION)

Due to the dynamic nature of seismic excitation, the actual displacements and stresses developed in a
structure are time dependent, i.e. they are functions of time (t). To analyze a structure under dynamic
loads, efficient methods have been developed that discretize and solve the model of the structure on the
basis of the Finite Element Method. Within this framework there exist methods that can perform a linear
or non-linear analysis, i.e. elastic, small deformation, or inelastic, large deformation analysis for a given
seismic excitation, expressed in the form of an accelerogram a(t). The cost of such analysis is generally
high, while the results correspond to a particular excitation and, as such do not offer a reliable basis for
design. To increase the reliability of the method, a set of artificial accelerograms that represent the
seismicity of a particular region is usually generated. This procedure, however, renders the method very
expensive.

Eurocode 8 [1] considers the use of time domain dynamic analysis, i.e. a direct dynamic analysis
performed by numerical integration of the differential equations of motion. It stipulates conditions for the
use of artificially generated accelerograms and discusses the overall reliability of the method. The
reliability must be at least the same as that obtained by the standard procedure of the Code which is the
response spectrum method. Although the direct dynamic methods can perform a close-to-reality analysis,
this approach is justified and can be employed effectively only for large and complex structures. It is used
where no previous experience of the structural behaviour exists, or for detailed evaluation of the response
of existing structures under specific earthquakes.

The cost of an analysis based on the finite element method can be kept reasonable by using only line
elements and by avoiding the use of surface elements. The mass of the structure of buildings is mainly
concentrated at the floor levels. This distribution permits the treatment of all the masses of the structure as
lumped at the floor levels in dynamic analysis. The dynamic degrees of freedom for which inertia forces
are developed can then be reduced to a reasonable number. All the remaining kinematical degrees of
freedom control the statics of the structure, and can then be expressed in terms of the dynamic degrees of
freedom. In this way the number of differential equations that express the dynamic response of the system
can be reduced to a small number, leading to reasonable and acceptable solutions.

3. RESPONSE SPECTRUM METHOD OF ANALYSIS

The time dependent solutions discussed above express the dynamic response of the structure due to a
particular earthquake given in the form of an accelerogram. They do not offer the required information for
design however, because one particular earthquake cannot be representative of the seismicity of the area
under consideration.

In order to define an envelope of different earthquakes and also to eliminate the factor of time, the
concept of the response spectrum was developed. The response spectrum provides the required
information for design purposes and, at the same time, simplifies the analysis by reducing the problem to
a static problem of the estimated maximum responses. The response spectrum is defined, on a single
degree of freedom system of varying frequency excited by a specific earthquake, as the maximum
response of the system, ignoring the particular time of its occurrence. If the response is the displacement
of the system then the displacement spectrum is formed. If the response is the velocity or the acceleration,
the velocity or acceleration spectra are developed. The acceleration response spectrum is of primary
interest in earthquake engineering. More details about earthquake response spectra are given in Lecture
17.2.

The response spectrum method of analysis is the standard design procedure of modern seismic design
codes, e.g. Eurocode 8. It aims to give directly the maximum effects of the earthquake in the various
elements of the structure.

The general method, called also the multi-modal method, consists of computing the various modes of
vibration of the structure and the magnitude of the maximum response in each mode with reference to a
response spectrum. A rule is then used to combine the responses of the different modes. For this reason
the method is also known as the superposition of modal responses method, although the same name is
used for linear dynamic analysis where the mode shapes are used to decouple the differential equations of
motion and convert the n-degree of freedom coupled system to n-single degree of freedom systems. The
combination rule will generally be a square root of the sum of squares (SRSS) of the various modal
responses. This combination rule must be applied to all computed quantities, i.e. bending moments, shear
forces, normal forces and displacements. As a consequence, the resulting internal forces do not represent
an equilibrated set. Where the frequencies of a structure do not differ by more than 10%, different
combination rules need to be employed. In Figure 1 the steps of such an analysis by means of the
response spectrum are briefly summarised.
The response spectrum method is valid only for linear behaviour of a structure, i.e. only for an elastic
analysis with small deformations. For this reason the term elastic response spectrum is generally used.
However an equivalent method can be developed which results from comparative linear and non-linear
analyses. It uses a modified response spectrum such that the output internal forces from a linear analysis
will be correlated with the non-linear ones. This modified spectrum is referred to as the design response
spectrum. It is derived from the elastic spectrum modified by factors that take into account the influence
of the non-linearity of the structural material, the soil and other damping characteristics. In Figure 2 the
design response spectra to be used in the analysis of structures, as given in Eurocode 8 [1], are shown
schematically.
The main advantage of using the design response spectrum is that the analysis is linear while the results
represent the non-linear response of the structure.

A more simplified procedure than the multi-modal method, is the so-called equivalent static force
analysis, sometimes also called, e.g. in Eurocode 8 [1], the simplified dynamic analysis. This method is a
particular application of the design response spectrum method where one particular mode of vibration is
predominant as compared to others. This is the case for regular buildings (regular stiffness and mass
distribution over the height of the building according to Eurocode rules, see Lecture 17.5). The system is
accurately modelled by a single degree of freedom system. In essence the design spectrum method is
reduced to one mode of vibration to express the dynamic behaviour of the system. Usually the first
flexural mode shape is considered as a primary mode of vibration which can be simplified further into a
simple line. The equivalent static forces are computed as shown in Figure 3. A classical static analysis can
then be performed under the action of these equivalent static forces. The only prerequisite of the method
is the fundamental period of vibration T of the structure. It needs to be calculated in order to find the
appropriate design spectrum value (T), necessary to compute the base shear V. Alternatively, if an
accurate value of the period T is not available, the value of the fundamental period can be calculated
approximately by using one of the recommended formulae.
The equivalent static force method is an approximate method which is adequate for certain types of
structures and for the preliminary design of other structures. There may be cases where this method is not
conservative because the contribution from higher modes of vibration may be significant. For these cases
a complete dynamic response spectrum analysis is advisable for the final design stage.

In Table 1 a summary of the possible methods of structural analysis under seismic actions is presented.
Moreover the following remarks can be made:
 The effects of earthquake on a structure depend upon its stiffness and mass characteristics. The
forces induced in flexible structures (high fundamental period T) are generally lower than those in
stiffer structures.
 The effects of earthquake on a structure depend upon the distribution of the mass and the stiffness
of the structure. Non-regular distribution involves the influence of more vibration modes on the
response.
 Simplified analysis methods, such as static equivalent force analysis, generally can be applied to
regular structures, but in some cases may give unsafe results.
 Non-regular structures require more sophisticated analysis, such as the response spectrum or
modal superposition method.
 Large complex structures with special features of behaviour should be analysed by more
elaborate methods such as non-linear dynamic analysis.
 The designer should always keep in mind that in all the above-mentioned methods of analysis,
many uncertainties have been rationalized. The control of the uncertainties requires compliance
with the rules of "good practice" mentioned in Lecture 17.5. The uncertainties relate to behaviour
of the structural material under cycling loading, discrepancy of the earthquake characteristics, real
damping factor, effects of soil-structure interaction etc.

It is clear from the above discussion that the design of an earthquake resistant structure is a complex task
which requires engineering judgement. It must be performed by experienced engineers. The blind use of
computer software as blackboxes may result in inadequate design.

4. INELASTIC BEHAVIOUR AND ITS ROLE IN DESIGN

The elastic design of an earthquake resistant structure leads to very expensive structures. Moreover it is
not consistent with the current design philosophy which seeks to establish controlled dissipative zones in
the structure where seismic energy can be dissipated by means of ductile hysteretic behaviour. The
principal dissipative zones in steel structures are plastic hinges (in bending), sheared web panels and
members under plastic tension (Figure 4).
In Figure 5 the difference in energy dissipation between the elastic and inelastic concept is presented. The
energy input Ei of an earthquake is counterbalanced inside the structure by the following sum of terms:

Ei = Ee + Ed + Eye + Ekin

where

Ee is the energy of elastic strain

Ed is the energy dissipated in a viscoelastic way

Eye is the energy dissipated by yielding

Ekin is the kinetic energy.


To obtain a stable earthquake resistant structure, either the energy input is minimized by means of special
techniques, such as base isolation of the building, or the dissipative terms in the right hand side of the
equation are increased. The term Eye must be increased as much as possible. It should be noted that by
taking into account elastoplastic energy dissipation, a considerable weight reduction of the structure is
achieved. In Figure 6 the moment rotation diagram of two equivalent beam elements is considered from
the point of view of energy dissipation. The resisting moment M1 required to resist an earthquake
elastically, is 3 times greater than the resisting moment M2 of the elastoplastic element with a ductility of
2. Expressed in terms of weight, beam 2 is only equivalent to 0,6 of beam 1. Thus the ductile behaviour
allows for substantial economy in the size of the elements of a structure. This economy is even more
substantial since the local ductility can be higher than 2. In steel structures the value of local ductility can
be as high as 10.
In order to design structures with dissipative behaviour by employing an elastic analysis which is easy for
the design office, certain rules have to be followed. They assure the safe formation of as many as possible
local dissipative zones, avoiding local failure mechanisms.

To approximate the results of a non-linear dynamic analysis by performing an elastic analysis, the
conventional response spectrum method is modified by reducing the spectrum in some way to account for
the inelastic energy dissipation of the real structure under the earthquake action.
This reduction is accomplished by using the structural behaviour factor q. It can generally be defined as
the ratio between the maximum accelerogram that a structure can withstand without failure and the
accelerogram for which yielding appears somewhere in the structure. The definition is general and can be
applied to different quantities of interest. In steel structures, one way to establish the correlation between
a conventional elastic analysis and the real inelastic behaviour is as follows:

For a given structure under a specific earthquake action a(t), a series of computations of the non-linear
dynamic response is performed by applying actions (t), where  is a multiplier. By increasing the value
of  the following successive situations emerge (Figure 7) [2, 3]:
  values are such that all sections of the structure remain elastic. In these cases, if d is a
displacement that characterizes the deformation of the structure, e.g. storey drift, then d will be
proportional to .

 The particular value of  which corresponds to the phase where yield stress is reached in one
section of the structure is called e.
 In the next phase, the  values are such that the real d's are smaller than the d's calculated by the
elastic analysis, i.e. supposing unlimited elastic behaviour, because of the energy dissipation by
yielding.
 By increasing the  values further, a max value is computed which corresponds to the same
elastic and inelastic displacement. This coincidence is due to the increasing role of P- effects,
which increase the displacements.

The behaviour factor q, is then defined as:

q = max /e

Thus the existence of a meeting point between the two forms of behaviour, allows a direct link between
the linear and non-linear computations. The equivalence states that, for a given accelerogram a(t) and a
known value of q, the usual linear analysis under the action a(t)/q and the usual checks on stresses, give
the same safety level as the dynamic non-linear calculations under the action of a(t). This equivalent is
due to the counteraction of the yielding effect which reduces the displacements, and the P- effect on the
structure which increases the displacements.

The real displacements of the structure ds are given as q times the elastic displacements de calculated by
using the reduced forces, i.e.

ds = q d e

The values of the factor q for various types of steel buildings are given in Lecture 17.5. All recent design
codes use a similar approach with slightly different values for the q factor. These discrepancies are
justified by the fact that q factors are not only functions of the shape of the structure, but they depend also
on the accelerograms a(t) considered. The accelerograms differ from one part of the world to the other.
Other points of difference may be due to the selected parameter characterizing the behaviour, which may
be the equal energy dissipation rather than the displacements, and due to the safety factors used for the
elastic analysis, which usually are higher than those used for the inelastic analysis. Thus the appropriate q
factors involve a theoretical approach but also an engineering judgement.

It should be noted also that the analysis using a q reduction factor for an earthquake action is
conventional. Safety in the various structural elements is assured by requiring the computed comparison
stresses to be less than or equal to the yield stress. For the design of connections, under a real earthquake,
the real comparison stresses are equal to fy in dissipative zones. It is for this reason that connections close
to dissipative zones must be designed to transmit the plastic design resistance of elements and not the
elastic internal forces computed on the basis of an elastic analysis using a q reduction factor.

5. CONCLUDING SUMMARY

 The design philosophy for structures to resist seismic actions requires that the structure
must not collapse and must retain its structural integrity under a "strong" earthquake. The
structure must also not be damaged or limited in use under a "moderate" earthquake. To
avoid collapse, the structure is allowed to develop plastic zones in which seismic energy
is dissipated.
 Methods given by modern design codes for the analysis of structures under seismic
actions assess their behaviour against these performance requirements.
 Time-domain methods are used but their application is expensive.
 The response spectrum method is the standard procedure of modern seismic design
codes, e.g. Eurocode 8. A simplified response spectrum method for regular buildings is
available.
 The elastic design of an earthquake resistant structure leads to very expensive structures.
Consequently the current design philosophy uses controlled dissipative zones in the
structure where seismic energy can be dissipated by means of ductile hysteretic
behaviour.

6. REFERENCES

[1] Eurocode 8: "Structures in Seismic Regions - Design", CEN, (in preparation).

[2] Ballio, G. (1985) ECCS Approach for the Design of Steel Structures to Resist Earthquakes.
Symposium on Steel in Buildings, Luxembourg. IASE-AIPC-IVBH Report Volume 48 pp 313-380.

[3] Ballio, G. (1990) European Approach to Design of Steel Structures. 1990, Proc of Hong Kong Fourth
World Congress - Tall Buildings: 2000 and Beyond, pp 935-946.

Table 1: Methods of analysis for structures under seismic actions

Data needed Type of analysis Use - Design Codes


DIRECT DYNAMIC ANALYSIS  Linear or non-linear  procedure permitted by
Codes but not for design
(Time domain)  Direct Integration
 Use only for large and
 Accelerogram a(t) complex structures

(real or artificial)  Use for evaluation of


response of existing structures
 Characteristics of the structure, under a specific earthquake
elastic & inelastic (e.g. M- curves for
connections)
RESPONSE SPECTRUM  Modal analysis (linear)  Standard design procedure in
ANALYSIS Seismic Codes
 Mode shape superposition
 Design Response Spectrum  No limitations of use

 Characteristics of the structure,


elastic only
EQUIVALENT STATIC FORCE  Static analysis  Procedure permitted by
ANALYSIS Codes for buildings with
 First vibration mode is specific limitations of
 Design Response Spectrum predominant regularity

 Characteristics of the structure,


elastic only

Previous | Next | Contents

Previous | Next | Contents

ESDEP WG 17

SEISMIC DESIGN

Lecture 17.5: Requirements and Verification

of Seismic Resistant Structures

OBJECTIVE/SCOPE

To present the general design principles and requirements for building structures in seismic zones.

PREREQUISITES

None.

RELATED LECTURES

None.

SUMMARY

The general principles (symmetry, regularity, redundancy, torsional resistance, diaphragms, ...) of an
earthquake resistant design are first discussed.

Complete details on structural design for steel buildings, based on the general principles and including
rules and checks of Eurocode 8 [1] are given. They include data on regularity, elements and connections,
typology of structures and the q factors, strength and ductility checks required for elements and
connections.

1. EUROCODE 8 - SAFETY VERIFICATIONS

Overview of the requirements

Designing a safe structure in earthquake regions is a multi-planar problem. The following table
summarizes the main requirements and criteria.
REQUIREMENTS CRITERIA
Ultimate limit states - checks on resistance, stability and ductility of
structural elements
No collapse under
- overall stability of structure
strong earthquake
- foundations
Serviceability limit states - checks on deformation conditions

Limitation of damage under

moderate earthquake
Other specific aseismic measures - planning and design

- height and other limitations

- foundations

- quality plan

- ground investigations

Ultimate Limit State

 Strength

For all structural elements, the design resistance Rd/Rd  design action effects Sd.

The resistance Rd is calculated according to rules specific to the material. Explanations are given in
Sections 3 and 4.

 Stability

Second order effects, are either taken into account explicitly, or they are checked as being negligible
using the following criterion (Figure 1).

M2nd order << M1st order

Ptot . dr << Vtot . h

=

where

Ptot is the total gravity load at and above the storey considered
dr is the design interstorey drift (dr = q . de !)

Vtot is the total seismic design shear at the storey considered

h is the storey height.

 Ductility

Checks on ductility are material related and are described in Sections 3 and 4.

 Foundations

The resistance of soil must satisfy "capacity design" requirements; this means that the foundations must
resist the maximum forces that the structural elements can transmit to them, regardless of the actual
values due to the seismic design actions.

Serviceability Limit State

Checks on deformation conditions

 Interstorey drift
For structures including non-structural elements sensitive to deformation, the interstory drift dr is limited,
e.g. 0,002h.

 Pounding

Joints between structures must be designed to avoid pounding between two adjacent structures.

2. GENERAL DESIGN CONSIDERATIONS FOR BUILDINGS IN EARTHQUAKE AREAS

Introduction

Some general principles for the design of structures to be erected in earthquake areas are given here. It
should be pointed out that earthquake resistant structures can be designed without consideration to these
principles.

Compliance with these principles will however substantially reduce the possibility of the occurrence of
dynamic effects which cannot be predicted by linear analysis. For this reason, Eurocode 8 [1] prescribes
lower values of seismic actions (higher q factors) for systems complying with the general rules. The
overcost of an earthquake resistant structure is reduced by use of these lower values in comparison with a
usual structure. It also seems that the combination "good design - simple analysis" gives safer structures
than the combination "bad design - refined analysis".

Principle 1 - Simplicity

The dynamic behaviour of a simple structure is easy to understand and to compute. The risk of forgetting
any special aspect of performance such as an interaction of parts with different rigidity is low. Overall
simplicity leads to simple detailing.

Principle 2 - Continuity and uniform distribution of strength

Any discontinuity in the design brings a stress concentration and, potentially, a local failure mechanism.
Energy dissipation in the structure should be as high as possible. There should therefore be many
dissipative zones in the structure. As a result a global failure mechanism should be aimed at. The non-
homogeneous behaviour of a structure with major discontinuity leads to tedious calculations and difficult
design of the connection areas.

Practical continuity has many aspects.

Detailing:

 There should be no weakening in sections.


 Secondary effects generated by offsets, as well as sudden changes in sections should be avoided.
 Connections should be away from dissipative zones.
 Site control should be effective to obtain a proper correspondence between design and execution.
Particular attention should be given to, for example, bolts, prestressing (minimum and maximum
yield strength, ductility of the material), no locking of the displacement of the structure by
unplanned infilled walls.
 There should always be positive links. Friction cannot be relied on to resist horizontal forces or
relative displacements of, for example, supports, diaphragms, girders of a bridge. Similarly,
gravity force is not enough to restrain non-structural elements. Disconnection of hanging ceilings
or claddings can be dramatic.

Overall design:

Redundancy is a minimum condition for developing real continuity in a structure. It is essential, but not
sufficient.

Continuity and uniform distribution of strength in the horizontal direction of a building generally means
symmetry, if possible almost axisymmetry. Plan layout of vertical resisting elements should also
recognise the need of a high global torsional stiffness. Major damage has been observed in the connection
zones of structures with 'wings'. The differences in flexural mode shapes of these 'wings' induce this
result, Figure 2.
Continuity in the vertical direction means a lack of setbacks and a relatively uniform distribution of the
shear and flexural resistance of the structure. The so-called "soft storey" should be avoided. Unintended
changes in rigidity caused by "non-structural" elements like infills, partition walls ... should also be
avoided, Figure 2. Eurocode 8 allows simplified methods of analysis of buildings when certain conditions
are met, see Table 1.

Table 1: Structural regularity in Eurocode 8

For the application of simplified methods of analysis, a building can be classified as regular when the
following conditions are satisfied simultaneously.

Geometrical and structural layout in plan

 The plan configuration does not present divided shapes nor large recesses. When re-entrant corners or
recesses exist their dimension does not exceed 25% of the building external dimension in the
corresponding direction.

 The structure of the building is distributed along an orthogonal mesh defining two main directions with
similar stiffnesses.

 The building has an approximately symmetrical plan configuration with respect to those two main
orthogonal directions.

 At any storey the distances (measured in the two main directions) between the centre of masses and the
centre of stiffness do not exceed 15% of the "resilience radius" defined as the square root of the ratio of
the storey torsional and translational stiffnesses.

 The in-plan stiffness of the floors is high enough, in comparison with that of the vertical structural
elements, such that a rigid behaviour may be assumed. Furthermore, the floors should not present large
holes hindering such assumption especially if they are located in the vicinity of the main vertical
structural elements.

Vertical configuration

 The stiffness and mass properties are approximately uniform throughout the building height.

 Where there is a gradual setback throughout the height, the setback at any floor is not greater than 20%
of the previous plan dimension in the direction of the setback and symmetry about the vertical axis is
preserved.

 If a setback greater than 20%, but not greater than 50% and preserving symmetry, occurs within the
lower 15% of the total height of the building above the surrounding ground level (or above the level of
application of the seismic excitation), it may still be classified as regular. In such cases the structure of the
base zone beneath a vertical projection of the upper storeys must be able to support at least 75% of the
shear forces that would develop in that zone in a similar building without the base enlargement.

 Where setbacks occur only in one facade, the overall setback (sum of setbacks at all storeys) is not
greater than 30% of the plan dimension in the first storey and at any floor the individual setback is not
greater than 10% of the previous plan dimension.

When circumstances, e.g. the available site, aesthetics or use of the building, are such that structural
continuity is not possible for the whole volume of the structure, the latter can be subdivided into smaller
blocks. Structural continuity can then exist in each block, the blocks being linked by flexible footpaths. A
proper distance computed as the sum of their maximum displacements must be left between two
contiguous blocks to avoid pounding of the blocks when they are excited by earthquake motion.

Principle 3 - Dissipative structures

Building structures able to dissipate energy are introduced and discussed in Lecture 17.4. Dissipative
zones must be safe and numerous. This situation can be achieved in different ways, by adopting the
design approaches based on the principles described below.

Principle 4 - Low slenderness

In general, the more slender a structure, the worse the overturning effect of an earthquake.

High slenderness may however be useful in some cases (see Principle 7).

Principle 5 - Torsional resistance

Earthquake action generates special torsional effects in structures, mainly because the resultant of inertia
forces generated by the earthquake is applied at the mass centre M of each floor of the structure and the
latter generally does not coincide with the torsion centre S of the earthquake resistant structure, Figure 3.
The resultant force times the distance to that centre gives a torsional moment Mt. In multi-storey frames,
the torsional moment from one particular floor is increased by the resulting moment of the floors above.
In most structures, the approach to evaluate this torsional moment is partly rational (the distance between
S and M) and partly statistical, because the load distribution in a structure is not well known at the design
stage and changes through the life of the structure. Codes indicate how to evaluate this second term. A
few structures are free of torsional effects (axisymmetrical), e.g. water towers.
There may also be a second cause for torsional action. The earthquake itself results mainly in the vertical
propagation of a shear wave so that two points of the structure may be moving differently at one time.
This origin of torsion is normally important for structures which are very large in plan, e.g. bridges.
To resist the torsional action, the structure must be given adequate torsional rigidity. The best solution is
obtained by putting the earthquake resistant part of structure close to the perimeter of the structure as a
whole and all around it, complying with the symmetry principle. It must be pointed out that the classical
"one vertical core" structure of earthquake free areas is not effective, because it lacks torsional rigidity. It
should simply be avoided in unsymmetrical layouts.

Principle 6 - Diaphragms

Diaphragms in a building are the structures which transfer horizontal inertia forces, resulting from the
motion applied to the masses of floors and their loading, towards the structures able to contain them.

Diaphragms must be structures of low deformability and capable of efficiently distributing the horizontal
action between the various vertical resistant structures. Diaphragms may be provided in many ways:
concrete slabs, composite slabs, trusses, frames. Diaphragms must be properly linked to the vertical
rigidity elements. The links must be able to transmit the horizontal inertia force.

Principle 7 - Rational distribution of loads in the structure

Important loads should not be put at places where they generate inertia forces under earthquake loading.
For example, a library should for preference be at ground level. An X-Ray installation should be close to
the centre of rotation. Masses should be reduced whenever possible. For instance, using light floor
systems rather than traditional slabs can bring drastic reductions in inertia forces and result in substantial
economy in the framework. Similar choices should be made for partition walls, infills, claddings, etc.

Principle 8 - Stiffness adapted to the site

The shape of the design response spectrum (Lecture 17.4) indicates that earthquake forces are lower for
structures characterized by a predominant high period (T) of vibration. This characteristic can sometimes
be used at the start of a design, especially when more refined data are available for a particular site. For
instance, in a site with thick alluvium layers, which is characterized by a response spectrum with
relatively high amplitudes in the high period range and low amplitudes in the low period range, a very
rigid structure would better fit than a flexible one. The opposite choice would apply to rock areas.

Principle 9 - A strict correspondence between the real structure and the model used in its analysis

Designing a structure which is safe under earthquake loading is feasible. However to achieve a safe
structure, the model used in the analysis must correspond to the real structure. Otherwise, for instance,
yielding will take place in other places than foreseen or will not take place and be replaced by a brittle
failure. In earthquake engineering more material or a stronger material does not mean more safety,
because safety is not only derived from strength, but also from ductility.

There are many causes for discrepancies between reality and model, for example:

 non-structural elements like infills must not give unexpected rigidity to a structure. Such rigidity
can completely change the behaviour of the structure, introduced high local shear and cause
failure. Non-structural elements must be linked in a way such that they do not in fact play any
structural role.
 distribution of yield strength throughout the structure should not differ much from that assumed,
otherwise yielding will take place elsewhere than foreseen or not take place.
 site control should ensure the real structure corresponds to that planned.
3. DESIGN OF STEEL STRUCTURES IN EARTHQUAKE AREAS

Materials

Materials such as structural sections, bolts and welds which are used for steel structures in earthquake
prone regions are not different from those used for steel structures elsewhere. They are usually submitted
to the same quality checks.

However, compliance to Principle 9 of Section 2 requires the definition by the designer of a maximum
value of yield strength of the steel to be used in the structure. This requirement is specific to earthquake
resistant design. The reason is that normally steel material is delivered on the basis of a guaranteed
minimum yield strength, but in practice it may have a far higher value of yield strength than that ordered.
This fact leads in general to conservative design which is not detrimental for normal steel structures, but
which can be harmful in the case of earthquake resistant steel structures. Overstrength effects in
dissipative parts of the structure can lead to a concentration of seismic energy dissipation at points where
it is not expected nor wanted, as for instance at the connections.

Therefore, for the dissipative parts of the structure both lower and upper values of yield strength are
specified in design and in ordering of the material. Moreover, sufficient control to avoid overstrength
must be undertaken through specific application rules.

General structural steels according to EN 10025 are used in earthquake resistant steel structures. Bolts
should preferably be high strength grades 8.8 and 10.9.

Sections

Steel sections in dissipative zones of the structure must be able to withstand yielding without significant
loss of bearing resistance. This requirement can be a problem in compressed parts of sections where early
local buckling can occur. To avoid local buckling, restrictions are placed on the width-to-thickness ration
b/t of the compressed flat parts of sections. These restrictions depend on the maximum intended overall
ductility of the structure. For this reason, steel sections are classified into three classes in accordance with
three levels of behaviour factor q, as indicated in the following table.

Behaviour factor q Section class

q<4 1

2<q4 2

q2 3

The limiting values b/t for the above three classes of section are given in Eurocode 3 [2].

An increase of ratio b/t results in a lower local ductility because of the appearance of local buckling. This
reduction therefore results in a reduction of the capacity of the structure to dissipate energy, which is
finally expressed by a smaller value of the behaviour factor q.

Connections
Connections should not be the location for failure, for the following reasons:

 their failure mechanism is generally not well known.


 they have low global ductility, because stress concentrations locally exhaust the available
ductility of the material.
 high strength bolts are not very ductile. In tension connections they may be subjected additionally
to prying forces which are also not well known.
 the heat affected zone close to welds is less ductile than the original material.

Therefore, a criterion is imposed according to which connections near dissipative zones of the structure
must have sufficient overstrength so that yielding occurs in the ductile members (overstrength criterion).

Welded connections made with full penetration butt welds are considered to satisfy the above criterion.

Welded connections made with fillet welds and bolted connections, in order to satisfy the above
overstrength criterion, must meet the following requirements:

Rd  1,20 Rfy

where

Rd is the design resistance of the connection

Rfy is the yielding resistance of the connected member.

The above condition can often be attained by an increase of the member section in the connection zone.
Figure 4 shows two bracing connections, where the fulfilment of the overstrength condition requires a
reinforcement of the connection zone either by a welded plate or by an additional bolted cleat.
In bolted connections, the failure of bolts in bearing must control the behaviour and not failure in shear.

From the above discussion it is evident that the overstrength condition can lead to expensive connections.
There are two possible ways to overcome this overstrength penalty:

 to use full penetration butt weld connections in dissipative zones.


 to reduce the member section and consequently the yielding resistance of the dissipative zone so
that the overstrength condition gives a less penalizing value of Rd.

Earthquake resistant structures - General considerations

The term "earthquake resistant structures" (ERS) refers to those structural systems of a building which are
designed to resist the horizontal seismic actions.

In dissipative steel ERS, i.e. structures which through inelastic hysteretic behaviour can be submitted to
considerable deformations without failure by dissipating large amounts of seismic energy, there are
essentially three structural systems used to resist horizontal seismic actions (Figure 5):

a. Moment resistant frames (MRF) or simple frames.

b. Concentrically braced frames (CBF) or concentric truss bracings.

c. Eccentrically braced frames (EBF) or eccentric truss bracings.


In general, frames are more flexible than braced truss structures. Therefore they may experience greater
horizontal displacements under equal seismic actions. Such displacements can be a problem with respect
to the "P- effect" under a strong earthquake or to "damage" under a moderate earthquake. Compliance
with the overstrength criterion may also be very expensive for members in bending.

Truss braced structure contrary to frames, are always stiff depending of course on their configuration.
Their capacity to dissipate seismic energy differs greatly from one type to the other. The ability of both
frames and truss structures to dissipate energy whilst resisting seismic action is quantified by the value of
the behaviour factor "q", which has been described in Lecture 17.4.

Figure 6 presents the values of q-factor for the various systems provided that regularity criteria are met. If
the building is not regular in elevation the listed q values shall be reduced by 20%. These values should
be considered as maximum allowable ones, even if in some cases direct dynamic non-linear analysis
indicates higher q values in the region of 10 or 12.
Earthquake resistant structures

Specific considerations - criteria according to Eurocode 8

Frames

Frames are structures that resist horizontal seismic actions mainly through bending of their members.
They have a large number of energy dissipative zones located near to the beam-to-column connections.
The energy is dissipated through cyclic bending behaviour.

During seismic design, it is assumed that the frame as a whole satisfies the basic criterion of avoiding the
creation of a soft storey.

Under this criterion, the aim is to form plastic hinges in the beams and not in the columns in a global
failure mechanism, except at the bases of the columns. This mechanism is the so-called "strong columns-
weak beams" concept (Figure 7). When the design is such that plastic hinges form in the beams rather
than in the columns, these hinges have the role of spreading yield through the structure. Moreover, the P-
 effect is reduced and interaction between axial force and biaxial bending moments in the columns is
avoided.

The concept of "strong columns-weak beams" is not applied to single storey frames, to the top floor of
multi-storey frames and at the base of columns where they are connected to foundations.

Beams

Beams are verified as having sufficient safety against lateral or lateral and torsional buckling failure.

To obtain safe plastic hinges in the beams, a check is made that the full plastic moment resistance and
rotation capacity are not decreased by compression and shear forces. To this end the following
inequalities are verified at the location where the formation of hinges is expected.
where

M and N are the seismic action effects taking account of the behaviour factor q

Mpd, Npd and Vpd are the ultimate resistances of the section at the plastic hinge

Vo is the shear force of the beam, considered as simply supported, due to vertical loads

VM = (MRA + MRB)/1 is the shear force due to the resisting moments MRA and MRB of the beam at its
extremities A and B, calculated with the upper value of yield strength.

Beam-to-column connections must satisfy the requirements for connections, considering the bending
resistance Mpd of the plastic hinge section, and shear force equal to (Vo + VM), as specified above.

Columns

Columns are verified in axial force and bending, the design values of bending moments MCD,c being
resistance design values, i.e. values derived from maximum design moments of column due to seismic
actions, multiplied by a suitable capacity amplification factor.

The most unfavourable shear force of the column due to seismic combination actions must respect the
following condition:

V/Vpd = < 0,5

The transmission of forces between beam flanges at a beam-column node is achieved by extending beam
flanges to stiffeners across the column.

Concentric truss bracings

General

In concentric truss bracings horizontal seismic forces are mainly resisted by members in axial loading
(tension or compression). In such systems ductile members are mainly the tension braces, because energy
dissipation in compression braces deteriorates quickly due to buckling. The usual types of concentric truss
bracings are the following:

Diagonal type

The alternating horizontal forces are resisted in this type by the corresponding tension braces only, while
the contribution of compression braces is neglected. The diagonal braces of alternating loading can be in
the same bay (X bracing) or in different bays of the same storey. In the latter case the quantity "Acos "
(where A is the area of the brace section and  is the slope to the horizontal) must not vary by more than
10% between two opposite braces in the same storey.

Type V or 

In this type both tension and compression braces are needed to resist the horizontal seismic forces (for
equilibrium reasons). The diagonal braces may have a V shape or a  shape, in which case they meet at
the middle of the upper beam without interrupting its continuity.

Type K

Bracings of this type, where the meeting point of diagonals intersects the column at an intermediate point,
do not offer any possibility of ductile behaviour because they demand the participation of the column in
the yielding mechanism. Therefore q = 1 for this type of bracing, and its use is not recommended.

Diagonals

Diagonals must be verified for the condition: N/Npd  1,0

where

N is the maximum tensile force due to seismic combination actions

Npd is the design resistance in tension

Satisfactory dissipative behaviour of diagonals depends on their slenderness. For this reason the following
condition must be satisfied:

=  1,5

where

is the effective slenderness of the diagonal

A is the cross section area

fy is the yield strength

Ncr is the ideal critical Euler load of the diagonal (= 2EI/12).

Note: The above condition  1,5 is equivalent to slenderness ratio   140 for steel Fe E 235,
and   114 for steel FE E 355.

Columns and beams


Columns and beams are capacity designed, i.e. they are verified for buckling under an axial load a cd N,
where N is the maximum axial load due to seismic combination actions and acd is a suitable amplification
factor.

In bracings of type V or  the horizontal beams are designed to resist their vertical loads, neglecting the
intermediate support provided by the diagonals.

Eccentric truss bracings

General

Eccentric truss bracings are a lateral load-resisting system for steel buildings which can be considered a
hybrid between conventional frames and concentric truss bracings. They combine most of individual
advantages of frames and concentric bracings, whilst they minimize their respective disadvantages. Figure
8 illustrates some common arrangements.

The main characteristic of eccentric truss bracings is that at least one end of every brace is connected in
such a way that the brace force is transmitted either to another brace or to a column through shear and
bending in a beam segment called a "link", denoted by the symbol 1s. Because shear and bending in the
link due to horizontal forces are of considerable magnitude, it is convenient to concentrate the ductility
requirements to that segment.

The most attractive feature of eccentric truss bracings for seismic-resistant design is their high stiffness
combined with excellent ductility and energy dissipation capacity.
The yielding mechanism of links depends on the ratio of 1s to the length 1o=2Mp/Vp where Mp and Vp are
the plastic strengths in bending and shear of the link. Theoretically if 1s/1o  1,0 the links yield in shear
(shear plastic hinge). However, experiments have shown that the effect of strain hardening is very
important and cannot be neglected. As a result, in order to assure the more desirable behaviour of links
that yield in shear, it is recommended that 1s/1o 0,8. When 1s/1o  1,3 the link yields in bending (moment
plastic hinges). Yielding of the link is mixed between the above two limits. In all cases there is a
possibility for adequate ductility.

Links are designed to provide enough ductility. The other members (bracings, columns and rest length of
beams) are capacity designed, so that yielding is confined to the links.

Links

Key elements in developing the full strength and rotation capacity of shear links are proper stiffening and
lateral bracing. Two-sided, full-depth stiffeners must be provided at the link end. Intermediate stiffeners
may be single sided for beam depths less than 600 mm, but are required on both sides of the web for
deeper beams.

The maximum distance between successive stiffeners is taken equal to 56tw-d/5 for 1s/1o  1,15 or equal
to 38tw - d/5 for 1s/1o  0,80. For intermediate values of 1s/1o a linear interpolation is made.

Lateral bracing must be provided at the link ends at the locations shown in Figure 9. Strong and stiff
lateral bracing at these locations is critical to the stability of both the link and the brace. A composite deck
by itself cannot be counted upon to provide adequate lateral support for the link ends. Transverse beams
are the preferred lateral-bracing system.
After the selection of the link section, all other truss members are designed to remain essentially elastic
under the forces generated by the fully yielded and strain-hardened link. This design requires an estimate
of the ultimate shear force that can be achieved by the link. The ultimate shear force should be taken at
least as:

Vult = 1,5 Vp

Columns and braces

Columns must be designed to remain essentially elastic under the ultimate link forces, as well as the
appropriate vertical load contributions.

Braces must not buckle. They are therefore designed for the axial forces generated by the ultimate link
shear given above. Experimental results show that ultimate link shear forces may sometimes exceed the
value of 1,5 Vp due to overstrength of the web or due to the presence of a thick composite concrete deck.
A conservative design of the bracings is therefore appropriate.

Diaphragms

The horizontal diaphragms and bracings must be able to transmit with sufficient overstrength the
earthquake forces to the various earthquake-resistant elements connected by them.

This condition is assumed to be fulfilled by using a magnification factor of 1,5 for the verification forces
obtained from the analysis. Eurocode 8 [1] also gives minimum detailing rules for diaphragms in
reinforced concrete.

Specific control measures

The details of connections, sizes and qualities of bolts and welds as well as the steel grade of the members
and the allowable maximum yield strength fy in the dissipative zones are indicated on the fabrication and
erection drawings.

At the different phases of fabrication and construction continuous checks are necessary in order:

 to guarantee that the specified maximum yield strength of steel is not exceeded by more than
10%.
 to guarantee that the distribution of yield strength throughout the structure does not substantially
differ from the distribution assumed in the design. This check aims at the achievement of
sufficient regularity in terms of yielding behaviour to prevent the energy dissipation from being
concentrated to one storey only (Figure 10).
 to guarantee that the stiffness and strength assumed in the design are not exceeded by more than
10%.
Whenever one of the above criteria is not fulfilled, either new computations of the structure and of its
details are made to demonstrate its efficiency, or changes are made to confer equivalent efficiency. For
instance, such a change could be the reduction of the member section so that its plastic resistance is equal
to the intended one (Figure 10). A change of this kind allows more reasonable dimensions of the
connection (end plates, bolts) since in the overstrength condition of connections Rfy is reduced because it
refers to the reduced section which becomes the dissipative zone.

Cross-section and boundary Stress distribution Class A Class B Class C


condition
(compression positive)

Rectangular hollow section Compression 33  37  41 

Tubular section Compression 50 2 70 2 85 2

Webs of I-profiles plastic elastic 66  78  90 

Webs of flanges of welded distribution distribution


sections
Compression 33  39  41 

Combined bending and


compression

plastic elastic

distribution distribution

Outstanding flanges of welded Compression 9 10  12 


box sections or flanges of I-
Profiles Combined bending and
compression

Combined bending and


compression

Flanges of Compression 20  22  26 

I-Profiles

General fy 235 275 355

 1 0,92 0,81
=

Table 2 Limit b/t ratio of compressed parts of cross-sections for different cross-sectional classes

4. CONCLUDING SUMMARY

 The main requirements for the design of a structure in earthquake regions are that it shall not
collapse under a strong earthquake and that damage is limited under a moderate earthquake.
 To meet these requirements design is based on general principles usually involving:

 simplicity

 continuity and uniform distribution of strength

 ability to dissipate energy

 avoidance of high slenderness

 torsional resistance

 stiffness adapted to the site

 correspondence between the real structure and the model used in its analysis.

Rules and checks are given in Eurocode 8 [1] based on these general principles covering materials,
sections, connections and the structural systems which provide earthquake resistance. Particularly
considerations relate to frames, beams, columns and truss bracings.

5. REFERENCES

[1] Eurocode 8: "Structures in Seismic Regions - Design", CEN (in preparation).

[2] Eurocode 3: "Design of Steel Structures": ENV 1993-1-1: Part 1.1: General rules and rules building,
CEN, 1992.

6. ADDITIONAL READING

1. ECCS-CECM-EKS: "European Recommendations for Steel Structures in Seismic Zones",


Technical Working Group 1.3: Seismic Design, N.54, 1988.
2. SEAOC: "Recommended Lateral Force Requirements and Commentary", 1990.
3. Popov, E. P. and Engelhardt, M. D., Seismic Eccentrically Braced Frames, USA.

Previous | Next | Contents

Previous | Next | Contents


ESDEP WG 17

SEISMIC DESIGN

Lecture 17.6: Special Topics

OBJECTIVE/SCOPE

To give an overview of specifications for the seismic analysis and design of special structures such as
bridges and storage tanks.

PREREQUISITES

Lectures 17: Seismic Design

RELATED LECTURES

None.

SUMMARY

The lecture is divided in two parts. Part 1 deals with bridges; Part 2 deals with storage tanks.

The design of bridges is addressed with specific reference to Eurocode 8, Part 2 [1]. The general concepts
for the analysis and safety checks of steel bridge structures are summarised and discussed. For tanks, a
comprehensive literature and code review is given, covering both the dynamic behaviour and the design
problems related to their seismic resistance.

1. BRIDGES

1.1 Introduction

Bridges have been heavily stricken during several past seismic events. In Japan significant damage was
caused by Kanto, Nankai, Fukui and Niigata (1964) events. Most of these failures were due to large
foundation settlements, which caused excessive relative displacements and, sometimes, rigid-body failure
of bridge decks due to lack of support.

The damages caused by the S.Fernando earthquake in 1971, was more closely related to the dynamic
behaviour of the bridge structures. The earthquake severely affected the Los Angeles highway system. In
this case failures were often due to poor performance of deck joints and bearings.

During the Loma Prieta earthquake of 1989, one section of the S.Francisco-Oakland Bay bridge and the
entire Cypress Street viaduct in Oakland collapsed. A heavy death toll and significant direct and indirect
economic losses resulted in the S.Francisco Bay Area.

The Bay Bridge failure was due to large longitudinal displacements of a deck section, which exceeded the
length of the beam supports causing rigid-body failure.
The Cypress Street overpass failed due to poor performance of columns. The failure was probably caused
by inadequate and poorly detailed horizontal ties providing insufficient confining action and shear
resistance.

1.2 General Guidelines and Basic Requirements

The philosophy for the seismic design of bridges is similar to that adopted for the design of building
structures, with the additional requirement that bridges should be serviceable after the design seismic
event. These structures are considered to be essential in the post-earthquake period to allow rescue and
aid teams to reach the stricken areas.

More precisely, non-collapse and serviceability (at least for emergency traffic) must be ensured for an
event (design event) having an acceptably low probability of exceedance during the design life of the
bridge. In addition only minor damage and no disruption of serviceability can be tolerated for seismic
actions with large probability of occurrence during the design life.

The requirements may be met by means of design rules which, according to Eurocode 8: Part 2 [1], can be
grouped into the following categories:

 Strength Verifications
 Ductility Verifications
 Capacity Design Verifications
 Control of Displacements and Connection Behaviour.

The aim of these verifications is to control the non-linear structural behaviour which, for economic
reasons, has to be relied upon during strong seismic motions. The aim is met by performing the following
fundamental design steps:

 definition of dissipative zones, e.g. plastic hinges in piers, where severe inelastic deformations
can be safely developed.
 verification of dissipative elements against design seismic actions (strength verifications).
 verification of the ductility of dissipative zones.
 verification of non-ductile elements, e.g. bearings, against actions resulting from resistance
design, i.e. the actions that ensure the hierarchy of resistances of the structural elements. This
verification is necessary to avoid brittle fracture modes and to allow for the development of
inelastic deformations in ductile members.
 verification of relative displacements in connections, to avoid rigid-body failure due to
unseating.

1.3 Seismic Actions

The input seismic action must include the following aspects:

 characterisation of motion in a "point", i.e. in a single supporting area.


 characterisation of the spatial variability of the motion, i.e. of the correlation among the seismic
inputs at the various supporting areas.

1.3.1 Motion at a point


A single component of motion can be described in terms of a response spectrum, a power spectrum, or a
time history representation.

In Eurocode 8: Part 2 [1], a site dependent response spectrum is defined for a horizontal component,
depending on peak ground acceleration, natural period and damping factor. The type of soil affects both
the shape and the intensity of the spectrum. For the vertical component the same spectrum must be
adopted, scaled by a factor of 0,7.

Alternatively a power spectrum or a set of accelerograms can be adopted, provided that they are
compatible with the site dependent response spectrum.

A simplified six-component input motion, including rotational excitations is also specified depending on
the horizontal response spectrum (or power spectrum) and on the ground S-wave velocity.

1.3.2 Spatial variability

The spatial variability of the input motion is important for long bridges. If the length of the structure is of
the same order of magnitude as the length of the relevant seismic waves, the usual hypothesis of the
seismic motion being equal and simultaneous at all supports points must be removed.

Stochastic models of the spatial correlation structure of the ground motion are now available. They are
based on theoretical considerations regarding the wave propagation mechanism as well as on strong-
motion data recorded by dense arrays of instrumentation. These models, usually given in terms of cross
spectral density functions, can be directly used, together with the single-point power spectrum, to perform
a random vibration analysis. Alternatively, a set of time-histories compatible with the space-time
correlation structure of the design input motion can be simulated. They are then used in step-by-step
linear or non-linear dynamic analyses.

Eurocode 8: Part 2 allows a simplified modelling and a response spectrum analysis taking account of the
spatial variability of ground motion [1].

1.4 Methods of Analysis

According to Eurocode 8: Part 2, different methods of structural analysis can be employed depending on
the deck stiffness and on the overall regularity of the bridge.

If the in-plane horizontal deck stiffness is very large compared with the flexural stiffness of the piers, a
simplified static analysis based on the hypothesis of a rigid deck can be adopted.

If the deck is not very stiff, but its flexibility can be adequately modelled by means of a single
deformation shape, a "fundamental mode" model, essentially based on the classical Rayleigh method, can
be adopted.

In other more general cases a full dynamic analysis and modelling must be performed.

A full dynamic model must be used also for skew bridges or when the bridge cannot be regarded as
regular with respect to length of spans or stiffness of piers.
A site-averaged response spectrum can be used to account, in a simplified way, for different soil
conditions at the various support points.

When the full dynamic model is used, a linear or non-linear analysis can be performed. Non-linear
analysis can be used, based on design spectrum compatible accelerograms, to assess ductility demands in
the dissipative elements and to verify that internal forces in non-ductile elements do not exceed elastic
limits.

Linear dynamic analysis can be performed by reducing the response spectrum ordinates by a factor
(behaviour factor or q-factor) which takes into account non-linear behaviour. The same reduced spectrum
(design spectrum) is used for the determination of the equivalent static forces to be introduced in the
fundamental mode and for rigid deck simplified analyses.

1.5 Non-Linear Behaviour and q-Factors

Different behaviour factors (q) are specified by Eurocode 8: Part 2, depending on the expected dynamic
behaviour in the non-linear range: the higher the expected ductility, the larger the values of q-factor [1].
The values range from 1 (no-ductility) for arch bridges to a value of 3,5 for high ductility bridges in
which most of the energy input is dissipated by the bending deformations of the piers.

The q-factor depends on both the type of structure and on the adopted detailing.

For bridges with steel piers, provisional values for q-factors are given below:

(1) Bridges with steel piers where the earthquake forces are resisted mainly by the piers:

Piers without bracing:

q=3

Piers with traditional bracing:

flexural failure q = 2

axial force failure q = 1

Piers with eccentric bracing:

q=4

(2) Bridges with steel piers where the input of earthquake energy is mainly dissipated at the abutments:

q = 1,2

For a vertical excitation analysis, a q-factor of 1 should always be taken.


1.6 Deck Bearings and Longitudinal Restraints

Deck bearings are not deemed to behave in a ductile way. For this reason, they must be generally checked
against capacity design actions. For example, bearing devices connecting a cantilever pier to the bridge
deck should be designed against the transverse shear which produces the ultimate bending moment at the
pier base and neglecting the inertia forces in the pier.

Particular attention must be devoted to the longitudinal vibrations of the deck, due to the following:

 Longitudinal oscillations may cause rigid body failure due to unseating in sliding (or rolling)
devices. This failure can be avoided by providing adequate supporting areas and/or introducing
links to restrain excessive displacements. Relative displacements, if evaluated by means of linear
dynamic analysis, must be multiplied by the q-factor value.
 Especially for the case of continuous-deck bridges, problems can arise in ensuring adequate
restraints to longitudinal oscillations. In this case, one of the abutments must carry all
longitudinal inertia forces of the deck. Dissipative restraint devices must be provided to avoid
excessive axial forces in the deck and to prevent large longitudinal displacements.

1.7 Provisions for Steel and Composite Bridges

According to Eurocode 8: Part 2, steel and composite bridges shall be designed in accordance with
Eurocode 3 [2] and Eurocode 4 [3]. The structure must then be verified under seismic conditions.

The ultimate limit states that must be considered in the design are the following:

 Failure of bearings due to combined shear and vertical forces.


 Excessive movement of the bearings which may lead to failure of transverse elements of the
superstructure or collapse of the head of the piers.
 Severe damage or failure of the piers, due also to P- effects.
 Severe damage (or failure) of the superstructure.

According to Eurocode 8: Part 2, seismic protection can be achieved by means of ductile behaviour of the
piers, or by introducing isolating devices between the superstructure and the piers [1].

Such devices must limit the transfer of excessive horizontal forces between the superstructure and the
piers, and should introduce additional damping.

The general concepts of dissipative elements and the rules of capacity design applied to prevent brittle
failures should be applied to steel bridges.

1.8 References

[1] Eurocode 8: "Structures in Seismic regions - Design" Part 2: Bridges, CEN (in preparation).

[2] Eurocode 3: "Design of Steel Structures": Part 2: Bridges, CEN (in preparation).

Part 1.1: General rules and rules for buildings, ENV1993-1-1, CEN, 1992.
[3] Eurocode 4: "Design of Composite Steel and Concrete Structures": Part1.1: General rules and rules for
buildings, ENV1994-1-1, (in press), Part 2: Bridges (in preparation).

1.9 Additional Reading

1. US Nuclear Regulatory Commission, "Seismic Input", Standard Review Plan 3.7, June 1975.
2. "Earthquake Resistance of Highway Bridges", Applied Technology Council, Palo Alto, California,
January 1979.
3. "Standard for Aseismic Resistant Design Specifications of Highway Bridges" by Japan Road
association for earthquake engineering, 1984.
4. "Guide Specifications for Seismic Design of Highway Bridges", American Association of State
Highway and Transportation Officials, 1983.

2. LIQUID STORAGE TANKS

2.1 Introduction

Tanks used as storage facilities for fluids ranging from non-toxic, inflammable liquids to highly toxic and
flammable chemicals are of special importance. Breakdown of the water supply (San Francisco, 1906),
uncontrolled fires which ignite adjacent tanks and buildings, and spillage (Niigata, 1964 and Miyagi-Ken-
Oki, 1978) or clouds of toxic chemicals may cause much more damage than the earthquake itself.
Jennings [1] gives a report on tank damage after the San Fernando earthquake. In a report by Wyllie et al
[2] tank damage caused by the 1985 Chile earthquake are described. Berz [3] summarises all major
natural disasters, including earthquakes, from 1960 to 1987. It is apparent that damage caused by
earthquakes plays a dominant role in the list of natural disasters. Further details of damage to oil refineries
from major earthquakes in the years 1933-1983 are given by Nielsen and Kiremdijan [4]. They conclude
that damage to the storage facilities of refineries has been severe and would indicate the need to upgrade
their performance.

Different modes of tank failure have been observed after earthquakes:

 elastic-plastic buckling of the tank wall near the bottom edge ("elephant footing"), caused by
axial compression forces due to the overturning moment, see Figure 1.
 elastic buckling of the tank wall, see Figure 2.
 elastic buckling due to low pressure near the top.
 failure of the roof (fixed or floating roof).
 failure of the bottom plate.
 failure of the foundation.
 fracture of piping at the connection to the tank.
 sliding of the tank.
The most relevant types of tank damage are elastic-plastic buckling ("elephant footing") and elastic
buckling of the tank wall.

To understand the behaviour of liquid storage tanks subjected to earthquake excitation, many research
groups have been active in investigating the dynamic behaviour of liquid filled shells. The aim has been
to develop methods for the earthquake resistant design of liquid storage tanks, and to provide codes for
engineers engaged in the construction of liquid storage tanks. Such engineering approaches need to be
based on consideration of the coupled dynamic system consisting of the elastic or elastic-plastic shell, the
tank's liquid content, and the deformable foundation. It is a liquid-structure-soil interaction problem.
Fundamental scientific findings were published by Housner [5], which allow estimation of the dynamic
loads of rigid tanks resting on rigid foundations. A recent review of the engineering treatment of storage
tanks under earthquake loading is presented by Rammerstorfer et al [6].

Recent engineering procedures, e.g. Fischer et al [7] and Veletsos and Tang [8], based on parametric
studies have been published. They allow an earthquake resistant design of storage tanks typical for the
petrochemical industry by simply using formulae and design charts.

There are many differences between current codes and recommendations for the earthquake resistant
design of liquid storage tanks. Nevertheless, global procedures are similar, especially for anchored tanks.
The design procedures can be divided into calculation of dynamically activated loads, and strength and
stability analyses. The following summary outlines the essentials of the procedures:

 Calculation of the dynamically activated loads by the application of the response spectrum
method:

 Calculation of the natural frequencies, damping values, and mode participation factors of the individual
vibration modes.

 Calculation of the maximum acceleration response of the individual vibration modes (rigid motion of the
"liquid column", interactive vibration of the flexible tank wall and liquid, sloshing vibration at the free
surface), see Figure 3.

 Calculation of the maximum contributions to the overturning moment (corresponding to the individual
vibration modes) due to the dynamically activated pressure caused by the horizontal earthquake
excitation.
 Superposition of the contributions to the overturning moment and the contributions to the dynamically
activated pressure caused by the horizontal excitation.

 Calculation and superposition of the contributions to the dynamically activated pressure caused by the
vertical earthquake excitation.

 Strength and stability analyses:

 Superposition of pressure caused by the horizontal earthquake excitation with that caused by the vertical
earthquake excitation with respect to different kinds of tank wall instability. Internal pressure is
stabilising with respect to elastic buckling and destabilising in case of plastic buckling.

For tanks not anchored to the ground, non-linearities arise due to the unilateral contact between the base
of the tank and its foundation in addition to non-linearities due to the elastic-plastic material behaviour.
To build a tank without anchors is much cheaper because neither a special concrete foundation nor special
anchors are needed. The dynamic behaviour of an unanchored tank is quite different from that of an
anchored tank. Partial uplift of the tank bottom caused by the overturning moment leads to increased
maximum axial compression forces in the tank wall. Hence, tank wall instabilities may occur at lower
overturning moments.

2.2 Anchored Tanks

2.2.1 Horizontal Earthquake Excitation

Housner [9] proposed a simple procedure for rigid tanks which is based on the response spectrum method.
Essential comments on the use of the response spectrum method for the calculation of the dynamic loads
of liquid storage tanks are given by Scharf [10].

In the 1970s it became clear that the influence of deformations of the tank wall, which is a thin shell, must
not be neglected and that the dynamic loads may be much higher than those of rigid tanks. Theoretical
considerations led to a simple model (Figure 4) used in current codes for the calculation of the dynamic
loads on earthquake - excited tanks in terms of the maximum overturning moment needed for the strength
and stability analyses. Summarising, the dynamically activated pressure acting on the tank wall due to a
horizontal excitation of a deformable cylindrical tank resting on a rigid soil is given by the superposition
of four pressure components:
PSL is the "convective" pressure component due to the fundamental sloshing vibration of the liquid
(circumferential wave number m=1).

PB is the "impulsive" pressure component due to the rigid body motion of the liquid which varies
synchronously with the horizontal ground acceleration.

PD is the pressure component due to the fundamental, i.e. m = 1, interaction vibration of the shell and the
liquid,

PD,m are the pressure components due to interaction vibrations with m  2. These components which result
from imperfections may be neglected with respect to the estimation of the overturning moment.

The maximum overturning moment is calculated by superposition of the individual contributions due to
sloshing, rigid body motion, and fluid-shell interaction vibration. Different superposition rules, based on
an SRSS superposition, have been proposed. In the above engineering approaches, formulae and diagrams
are presented for estimating the natural frequencies and the individual masses and height in Figure 4
which depend mainly on  = H/R, a size parameter R or H, and the mass density of the liquid. These
formulae and diagrams result from integration of the individual pressure contributions.

On this basis the maximum overturning moment resulting from the dynamically activated pressures acting
on the tank wall (bottom pressure not included) can, according to Fischer et al [7], be estimated by:

MM = [(MSL ASL HSL)2 + (MB AB HB)2 + (MD AD HD)2]1/2 (1)

or alternatively, according to Haroun and Housner [11] by:

MM = [(MSL ASL HSL)2 + (MB AB HB + MD AD HD)2]1/2 (2)

where, for tanks designed according to DIN 4119, the ratio between effective masses (MSL, MB, MD,
compare Figure 4) and the mass of the liquid content (MT) can be taken from Figure 5, and the
corresponding heights from Figure 6.
ASL, AB and AD are the effective accelerations. They are the product of the corresponding spectral
accelerations (taken from the response spectrum) and mode participation factors. The fundamental
frequencies can, according to Fischer et al [7] be approximated by:

fSL = [1/(2)] [1,84 g tanh (1,84 )/R]1/2 cps (3)


fD = [E s1/3/(LH)]1/2 / (2Fs ()R)] cps (4)

with Fs = 0,157 2 +  + 1,49 ;  = H/R (5)

and where

s1/3 is the wall thickness at H/3.

The pressure amplitudes at the bottom edge - values needed for the stability check and strength
assessment of the tank wall - can be estimated by Figure 7.

The two-dimensional character of the horizontal earthquake action is discussed by Scharf [10]. It is shown
that consideration of the one-dimensional acceleration is, in general, not sufficient. Relevant procedures
are presented by Scharf [10].

2.2.2 Vertical Earthquake Excitation

The vertical component of the earthquake leads mainly to the excitation of axisymmetrical vibration
modes. Again rigid body motion, interaction vibration of flexible shell and liquid as well as free surface
sloshing, can be distinguished, see Fischer et al [7]. They can be estimated by simple formulae similar to
the procedure described above, taking into account the radiation, i.e. geometrical, damping due to energy
radiation by outrunning waves, see Seeber [12].
2.2.3 Stability and Strength Analysis

From these estimates, the strength analysis and the buckling check can be performed. The empirical
formula developed by Rotter and Seide [13] for cylindrical shells under axial compression and internal
pressure:

nxcrit = 0,605 (Es2/R) [1 - (pR/(sy))2] *

* [1 - 1/(1,12 + k1,15] (y/250 + k)/(1 + k),

where nxcrit is the critical axial membrane force

y is the yield stress

p is the internal pressure

E is Young's modulus

k is R/(400s)

R is the tank radius

s is the thickness of tank wall (at the bottom)

leads to proper estimates. This result can be seen, for example in Figure 8 for the Friuli 1976 earthquake,
for case of steel tanks for the petrochemical industry.
2.3 Unanchored Tanks

In the analysis of unanchored tanks, the investigation of the retaining action of the bottom plate is
essential to obtain the increased axial membrane compression force in the tank wall.

Since a strongly non-linear (geometrical and material) fluid-structure-soil interaction and contact problem
has to be solved, the calculation of the dynamic response of unanchored tanks is very complicated. Hence,
no fully satisfactory models are available.

Clough [14] assumed that the uplifted tank rests, on the one hand, on a section of the circumference and,
on the other hand, on the area of an eccentrically positioned circle, see Figure 9a.
Wozniak and Mitchell [15] presented an improved uplift model taking the formation of plastic hinges in
account, see Figure 9b.

A more sophisticated model is based on the "shell-spring model" for uplifting strips by Auli et al [16].
Friction between the bottom plate and the soil, as well as membrane forces in the bottom plate, were
considered. The validity of the model was verified by experiments using Mylar model tanks, see
Rammerstorfer et al [17].

Results of comprehensive numerical analyses, Scharf [10], show a strong influence of the roof or top
stiffness on the axial membrane compression force distribution. This fact has also been reported by
Natsiavas [18] and Sakai et al [19]. Figure 10 shows the axial membrane force nx at the tank bottom for a
tank with a low and a high top stiffness (roof or edge ring) at different earthquake intensities. For the tank
with a high top stiffness, the maximum of the axial compression force lies on the axis of symmetry. For
the tank with a low top stiffness two maxima exist, which lie beside the axis of symmetry. This
phenomenon is not considered in most analytical models described above.

From results of parametric studies, a design chart has been developed which allows the estimation of the
maximum axial membrane compression force at the tank bottom of unanchored tanks N unanch, Scharf [10].
The chart is based on the maximum axial membrane compression force for anchored tanks Nanch (see
Figure 11). It is concluded that the increase of the axial membrane compression force must not be
neglected, especially in the case of "tall" tanks. The influence of the increased axial membrane
compression force on different kinds of instabilities of tank walls is remarkable.
A procedure in which the non-linear character of the vibration of uplifting tanks is considered in Scharf
[10] and Fischer et al [7].

2.4 Current Design Codes and Recommendations

Some of the existing codes and recommendations are listed below:

2.4.1 American Codes

 US Atomic Energy Commission, ERDA TID 7024, Nuclear Reactors and Earthquakes, 1963.
Derived from the proposals of Housner [5]. Recent findings are not included.
 API Standard 650, Welded Steel Tanks for Oil Storage, 1988. This rather old code is based on the
work of Wozniak and Mitchell [15] with respect to uplift.
 American Water Works Association, Standard D 100-84, AWWA Standards for Welded Steel
Tanks for Water Storage, 1984. This code applies to water storage tanks. Many effects with
respect to earthquake loading are not taken into account.
 American Society of Civil Engineering, Guidelines for the Seismic Design of Oil and Gas Pipeline
Systems, 1984. This recommendation was prepared by A S Veletsos and gives a comprehensive
state-of-the-art report. Flexibility of the tank wall is considered, but the procedures for
unanchored tanks might be improved.

The procedures of TID 7024, API 650 and AWWA-D100-84 are compared by Bureau [20]. The codes
were found to underestimate or to overestimate the dynamic response, depending on the distance from
and magnitude of the earthquake.

2.4.2 Austrian Recommendations


 Fischer, F.D., Rammersorfer, F.G., Scharf, K., Earthquake Resistant Design of Anchored and
Unanchored Liquid Storage Tanks Under Three-Dimensional Earthquake Excitation, 1990.

This report gives a summary of the Austrian research project which deals with the earthquake resistant
design of tanks. The scientific background of a design procedure is presented by the authors [7] and
considers:

 flexibility of the tank wall.


 three-dimensional earthquake excitation.
 the increased axial membrane force of uplifting tanks.
 the decrease of the natural frequencies due to uplift.
 different instability modes for the tank wall.

2.4.3 Canadian codes

 CSA Z276-M1981, Liquid Natural Gas (LNG)-Production, Storage and Handling, 1981. Many
effects of earthquake excitation are not taken into account. Only qualitative recommendations
are given.

2.4.4 Japanese codes

 Institute of Industrial Science, University of Tokyo, Draft of Anti-Earthquake Design Code for
High-Pressure Gas Manufacturing Facilities, 1981.
 Ministry of International Trade and Industry, Standard of Seismic Design for High Pressure Gas
Facilities, 1981.
 Fire Defense Agency of the Ministry of Home Affairs, Notification Specifying the Details of
Technical Standard on the Regulations of Dangerous Objects, 1983.

Wall deformations are considered in Japanese codes. Some recommendations for dealing with buckling
are given. Uplift of the tank bottom edge is not taken into account adequately for unanchored tanks.

2.4.5 New Zealand Codes

 Priestley et al, Seismic Design of Storage Tanks, 1986. These recommendations are mentioned
especially since they are a very comprehensive and well-formulated code reflecting all research
results up to 1985. In addition rectangular tanks and cylindrical tanks with axis horizontal are
dealt with.

3. CONCLUDING SUMMARY

 The philosophy for the seismic design of bridges is similar to that adopted for the design of
building structures, with the additional requirement that bridges should be serviceable after the
design seismic event.
 The requirements for bridges may be met by means of design rules according to Eurocode 8:
Part 2. The design steps required include definition of the seismic input and dissipative zones
and verifications of dissipative elements for strength and ductility and of non-ductile elements
for resistance and connections for relative displacements.
 For tanks, several different modes of failure may occur in earthquakes. The most relevant
modes are elastic-plastic buckling and elastic buckling of the tank wall.
 The design of anchored tanks considers horizontal and vertical earthquake excitation, and
includes stability and strength analysis.
 For unanchored tanks calculation of the dynamic response is very complicated and no fully
satisfactory model is available, but several procedures have been developed.
 Several design codes and recommendations include guidance for storage tanks.

4. REFERENCES

[1] Jennings, P.E. (Ed), "Engineering Features of the San Fernando Earthquake", EERI-71-02, pp. 434-
470, California Institute of Technology, Pasadena, 1971.

[2] Wyllie, L.A., Bolt, B., Durkin, M.E., Gates, J.H., McCormick, D., Smith, P.D., Abrahamson, N.,
Castro, G., Escalante, L., Luft, R., Olson, R.S. and Vallenas, J., "The Chile Earthquake of March 3,
1985", Earthquake Spectra, Vol.2, No. 2, Chapter 5, pp. 373-409, 1986.

[3] Berz, G. "List of Major Natural Disasters, 1960-1987", Natural Hazards, Vol. 1, pp. 97-99, 1988.

[4] Nielsen, R., and Kiremidjian, A.S., "Damage to Oil Refineries from Major Earthquakes", Journal of
Structural Engineering, ASCE, Vol. 112, pp. 1481-1491, 1986.

[5] Housner, G.W., "The Dynamic Behaviour of Water Tanks", Bulletin of the Seismological Society of
America, Vol. 53, pp. 381-387, 1963.

[6] Rammerstorfer, F.G., Scharf, K. and Fischer, F.D., "Storage Tanks Under Earthquake Loading", Appl.
Mech. Rev., Vol. 43, pp. 261-282, 1990.

[7] Fischer, F.D., Rammerstorfer, F.G. and Scharf, K., "Earthquake Resistant Design of Anchored and
Unanchored Liquid Storage Tanks Under Three-Dimensional Earthquake Excitation", Structural
Dynamics - Recent Advances, Schueller, G.I. (Ed), Chapter 5.1, pp. 317-371, Springer-Verlag, 1991.

[8] Veletsos, A.S., and Tang, Y., "Soil-Structure Interaction Effects for Vertically Excited Tanks",
Proceedings of the 9th World Conference on Earthquake Engineering 9WCEE, Tokyo/Kyoto, Japan, Vol.
VI, pp. 631-636, 1988.

[9] Housner, G.W., "Dynamic Pressure on Accelerated Fluid Containers", Bulletin of the Seismological
Society of America, Vol. 47, No. 1, pp. 15-35, 1957.

[10] Scharf, K., "Beiträge zur Erfassung des Verhaltens von erdebenerregten, oberirdischen
Tankbauwerken", Doctoral Thesis, Fortschritt-Berichte VDI, Reihe 4, Nr 97, VDI Verlag, Düsseldorf,
FRG, 1990.

[11] Haroun, M.A. and Housner, G.W., "Earthquake Response of Deformable Liquid Storage Tanks",
Journal of Applied Mechanics, ASME, Vol. 48, pp. 411-417, 1981.

[12] Seeber, R., "Das dynamische Verhalten fernerregter flüssigkeitsgefüllter Tankbauwerke auf
elastischem Untergrund, Doctoral Thesis, Institute of Mechanics, University of Mining and Metallurgy,
Leoben, Austria, 1988.
[13] Rotter, J.M. and Seide, P., "On the Design of Unstiffened Shells Subjected to an Axial Load and
Internal Pressure, Proceedings of the ECCS Colloquium on Stability of Plate and Shell Structures, Ghent
University, Belgium, pp. 539-548, 1987.

[14] Clough, D.P., "ExperimentaL Evaluation of Seismic Design Methods for Broad Cylindrical Tanks",
UCB/EERC-77/10, University of California, Berkely, 1977.

[15] Wozniak, R.S. and Mitchell, W.W., "Basis of Seismic Design Provisions for Welded Steel Oil
Storage Tanks", Proceedings of the Session of Advances in Storage Tank Design, API Refining Dept., pp.
485-493, 1978.

[16] Auli, W., Fischer, F.D. and Rammerstorfer, F.G., "Uplifting of Earthquake-Loaded Liquid-Filled
Tanks", Proceedings of the Pressure Vessels and Piping Conference, ASME, PVP Vol. 98-7, pp. 71-85,
1985.

[17] Rammerstorfer, F.G., Billinger, W. and Fischer, F.D., "Stabilität flüssigkeitsgefüllter unverankerter
Zylinderschalen auf schräger Unterlage, Zeitschrift für angewandte Mathematik und Mechanik ZAMM",
Bd 68, T240-T243, 1988.

[18] Natsiavas, S., "Simplified Models for the Dynamic Response of Tall Unanchored Liquid
Containers", Proceedings of the Pressure Vessels and Piping Conference, ASME, PVP Vol. 157, pp. 15-
21, 1989.

[19] Sakai, F., Isoe, A., Hirakawa, H. and Mentani, Y., "Experimental Study on Uplifting Behaviour of
Flat-based Liquid Storage Tanks Without Anchors", Proceedings of the 9th World Conference on
Earthquake Engineering 9WCEE, Tokyo/Kyoto, Japan, Vol. VI, pp. 649-654, 1988.

[20] Bureau, G., "Seismic Design Guidelines for Liquid Storage Tanks: Applicability and Limitations",
Proceedings of the 4th International Conference on Soil Dynamics and Earthquake Engineering, Mexico,
City, Mexico, pp. 343-354, 1989.

[21] Niwa, A. and Clough, R.W., "Buckling of Cylindrical Liquid-Storage Tanks Under Earthquake
Loading", Earthquake Engineering and Structural Dynamics, Vol. 10, pp. 107-122, 1982.

Previous | Next | Contents

You might also like