You are on page 1of 119

EUROCODE 8 – DESIGN OF STRUCTURES FOR

EARTHQUAKE RESISTANCE
EUROKODE 8 - PROSJEKTERING AV KONSTRUKSJONER FOR
SEISMISK PÅVIRKNING

A complementary
Guide Book to the seminar
Volume 1.0 (English)

Editor:
Dominik H. Lang
NORSAR, Kjeller (Norway)

April 2010

1
Eurocode 8 – Design of Structures for Earthquake
Resistance
A complementary Guide Book to the seminar
Volume 1.0 (English) – preliminary version –
Dominik H. Lang (ed.)
118 pages
Kjeller, Norway
April 2010
EC8 Guide Book Table of contents

Table of Contents
1 Ground conditions and seismic action ......................................................................... 7
1.1 Introduction ......................................................................................................................................... 7
1.2 Identification and classification of ground type ............................................................................. 8
1.3 Soil amplification............................................................................................................................... 16
1.4 Site-specific amplification factor..................................................................................................... 17
1.5 Topographic amplification .............................................................................................................. 20
1.6 Representation of seismic action .................................................................................................... 22
2 Seismic design principles............................................................................................ 27
2.1 Basic principles of conceptual design ............................................................................................ 27
2.2 Primary and secondary seismic members ...................................................................................... 34
2.3 Criteria for structural regularity ....................................................................................................... 34
3 Soil-structure interaction............................................................................................. 41
3.1 Introduction ....................................................................................................................................... 41
3.2 Code approach to Soil Structure Interaction analyses ................................................................. 41
3.3 Basic concepts of SSI ....................................................................................................................... 42
3.4 Step procedure to include SSI ......................................................................................................... 45
3.5 Kinematic effects .............................................................................................................................. 46
3.6 Foundation Damping ....................................................................................................................... 47
3.7 Analysis of impedance functions .................................................................................................... 48
3.8 Evaluation of seismic loads ............................................................................................................. 53
3.9 Stiffness and capacity of pile foundation....................................................................................... 54
3.10 SSI effects on the seismic response of structures ........................................................................ 54
4 Structural analysis ....................................................................................................... 57
4.1 Importance classes and importance factors .................................................................................. 57
4.2 Methods of analysis proposed by EC8 .......................................................................................... 58
5 Design of concrete buildings ...................................................................................... 81
5.1 Tutorial 5.1: Building based on precast concrete elements......................................................... 81
6 Geotechnical aspects and foundations ....................................................................... 87
6.1 Introduction ....................................................................................................................................... 87
6.2 Geotechnical earthquake engineering ............................................................................................ 87
6.3 Important common definitions ...................................................................................................... 88
6.4 Ultimate limit state (ULS) and damage limitation state (DLS) ................................................... 88
6.5 Dynamic soil properties ................................................................................................................... 88
6.6 Material factors .................................................................................................................................. 94
6.7 Slope stability ..................................................................................................................................... 95
6.8 Liquefaction ....................................................................................................................................... 96
6.9 Foundation system ............................................................................................................................ 98
6.10 Shallow Foundations ...................................................................................................................... 101
6.11 Pile Foundations ............................................................................................................................. 110

3
4
EC8 Guide Book Introduction

Preliminary remarks

The guide book at hand provides useful information in addition to the provisions given in Eurocode
8 – Part 1 (EN 1998:1) and partly to its more specific regulations. The guide book was developed by
all contributors to the Eurocode 8 seminar conducted by NORSAR, Standard Norge and NJF.
Its main purpose is to provide the participants of the course with background information on the
various topics, to guide them through the code and to corroborate specific problems with examples
and tutorials.

Dominik H. Lang
NORSAR, Kjeller

5
6
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

1 Ground conditions and seismic action


Dominik H. Lang (NORSAR, Kjeller) and Farzin Shahrokhi (Multiconsult, Oslo)

1.1 Introduction
The local soil profile at a construction site can have a profound effect on earthquake ground
motions. Local soil conditions can affect the amplitude, frequency content, and duration of strong
shaking.
The influence of local soil conditions on seismic ground motions can be assessed either based on
code-prescribed response spectra for soil sites (EN1998-1:2004, 3.1.2, Tables NA.3.1 and NA.3.3) or
by conducting a site-specific seismic ground response analysis.
Site amplification factor for seismic design in EC8 provision is generally based upon assumptions
summarized as:
- vertically propagating shear wave through uniform soil layers,
- the soil layers are horizontal and extent infinite laterally (level ground),
- the ground surface is horizontal,
- the bedrock surface is horizontal.
General curves such that are shown in Figure NA.3(903) of EN1998-1:2004 are based upon
approximate studies. When such data are used in building codes to account for soil amplification, the
most conservative assumption about the amplification potential of the soils is assumed.
The influence of vertical motions, compressional waves, laterally non-uniform soil conditions,
incoherence and spatial variation of ground motions because subsurface topography of bedrock
(basin effects) are typically not accounted for in code-prescribed amplification factor S.
The choice of the approach to employ is usually a decision of the design engineer and depends on a
variety of factors, including local soil conditions, surface topography and topography of subsurface
bedrock, the type of facility, and the importance of the project. For the following state a site-specific
ground response analysis is usually required:
- for major projects and critical facilities, when an analysis more accurate than a gross
empirical analysis is desired,
- for deep deposits of the soft clay and other special study soil sites (i.e. S1 and S2 as defined in
EN1998-1:2004),
- when the topography of subsurface bedrock shows a dramatic variation under the
construction site
- when the soil layers are not horizontal or are changing laterally.
Number 1 or 2 will require a 1D site response analysis, however condition 3 or 4 dictate a 2D/3D
site response analysis (choosing the equivalent linear or nonlinear approaches for site response

7
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

analysis will be based on local site conditions, i.e. if the soil layer(s) are exposed to liquefaction or
high strain degradation).

1.2 Identification and classification of ground type


The influence of the local soil condition on the seismic response of structures can be quantified by
defining ground types with different mechanical properties. Ground types A, B, C, D and E have
been selected to identify the soil profiles. Table NA.3.1 of EN1998-1:2004 provides for each ground
type a description of the stratigraphic profile and the parameters used to classify the soil. Three
parameters are used:
- average shear wave velocity of the upper 30 m of the soil profile vs,30 ,
- number of blows evaluated with the Standard Penetration Test NSPT ,
- undrained cohesive resistance cu .
It must be noted that the site classification should preferably be based on the vs,30 if available. SPT
tests are not common in Norway, therefore clause NA3.1.2(1)(901) of EN1998-1:2004 suggests to
use the CPT test results instead.
Two special ground types, namely S1 and S2, have been listed in Table NA.3.1 of EN1998-1:2004.
Ground type S1 includes deposits consisting (or containing) a layer at least 10 m thick of soft
clays/silts with high plasticity index (PI > 40) and high water content. Ground type S2 accounts for
all other soil profiles, and includes deposits of liquefiable soils and sensitive clay. Soil type S2 is thus
likely to fail under earthquake ground motion, which may cause severe structural damage. Clause
3.1.2(4) requires special studies for ground type S2 .
Similarly, soil type S1 can generally produce anomalous seismic site amplification and soil–structure
interaction effects, which significantly influence the earthquake characteristics hence seismic action at
the construction site. Soil S1 exhibits, in fact, very low values of shear wave velocity, low internal
damping and an abnormal range of linear behavior. Liquefaction of soil leads to catastrophic failure.
On leveled ground, it generally causes loss of bearing capacity of foundation systems, and on sloping
ground it gives rise to flow conditions, although on very gentle slopes lateral spreading occurs.
Detailed geotechnical studies should assess the effects of the thickness of soil layers and shear wave
velocity of the soft clay/silt layer and variation of stiffness between layers of soil S1 and underlying
materials on the response spectrum.
Using only the ground types defined in Table NA.3.1 of EN1998-1:2004 to assess the stratigraphy at
a given construction site may result in extreme oversimplification. Consequently, further
classification of the ground conditions can be made to conform more closely to the stratigraphy of
the site and its deeper geology.
There are two important comments which should be noted in calculating the average shear wave
velocity to classify the site soil:
1. The average shear wave velocity for top 30 m is considered from the ground surface and not
from foundation embedment depth (compare to Tutorial 1.2).
2. If the depth to bedrock is smaller than 30 m, the bedrock should not be included in
calculating the average shear wave velocity, and the equation (3.1) in clause 3.1.1.2 would
change to (compare to Tutorial 1.3):

8
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

h i
v s ,average  i 1
N
in [m/s] eq.(1.1)
hi

i 1 v s ,i

Figure 1.1. Illustration of exemplary soil profiles for ground types A – E.

1.2.1 Shear wave velocity vs


Shear wave velocity vs is a parameter that is directly related to the dynamic properties of the material
(here: the soil).

G
vs  in [m/s] eq. (1.2)

with: G – shear modulus in [kN/m2]
 – material density in [t/m3]

Tutorial 1.1: Calculation of shear wave velocity

Sand, densely bedded Rock


G = 120 MN/m2 = 120,000 kN/m2 G = 5,170 MN/m2 = 5,170,000 kN/m2
 = 1900 kg/m3 = 1.90 t/m3  = 2,300 kg/m3 = 2.30 t/m3
G 120,000 G 5,170,000
vs    251 m/s vs    1,500 m/s
 1.90  2.30

9
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

Shear wave velocity values can be derived form a number of experimental (seismic) methods
- cross-hole method
- down-hole method,
- surface wave reflection or refraction,
- suspension logging,
- spectral analysis of surface waves (SASW),
- Reflection microtremor (ReMi).

For soil materials the compression wave velocity vp exceeds the shear wave velocity vs by an amount
that depends on the compressibility (reflected in Poisson’s ratio ) of the solid (Kramer, 1996). The
ratio between vp and vs is described by equation (1.3):

vp ( 2  2  )
 eq. (1.3)
vs (1  2  )

Poisson’s ratio is a dimensionless parameter that represents the load-dependent ratio between the
specimen alterations in transversal and longitudinal direction. The values of Poisson’s ratio for soil
materials vary between 0.2 (rock) and  0.5 (water-saturated loose sediments).

1.2.2 Average shear wave velocity vs,30


Acc. to EN1998-1:2004, clause 3.1.2(2) a site should be preferably classified according to the value
of the average shear wave velocity vs,30 , if this is available. Otherwise the value of average NSPT should
be used. In contrast to many sources where the calculation of average shear wave velocity is
suggested in a erroneous way, the only way to compute average shear wave velocity (here: of the
uppermost 30 m) is done in the following way (EN1998-1:2004, clause 3.1.2(3), eq. (3.1)):
30
v s , 30  N
in [m/s] eq. (1.4)
hi

i 1 v s ,i

with: hi – thickness of layer i in [m]


vs,i – shear wave velocity of layer i in [m/s]

10
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

Tutorial 1.2: Classification of a velocity profile

Layer i hi [m] vs [m/s]


1 3.7 157
2 4.8 320
3 13.4 279
4 5.8 468
5 7.7 362
6 9.6 412

Calculation of average vs of the uppermost 30 m:


30 30
v s , 30    285 m/s
h1 h2 h3 h4 h5 *
3 . 7 4 .8 13 .4 5.8 2.3
       
v s ,1 v s , 2 v s , 3 v s , 4 v s , 5 157 320 279 468 362

→ the site would be classified into Ground type C (180  vs,30  360 m/s)

11
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

Tutorial 1.3: Classification of a site with shallow sedimentary layers (H < 30 m)

Layer i hi [m] vs [m/s]


1 5.0 375
2 5.0 646
3 10.0 1030
4 30.0 2126
5 > 80.0 2300

Problem: Average shear wave velocity vs of the uppermost 30 m clearly classifies the
site as a rock site (class A; vs,30 > 800 m/s)
30 30
v s , 30    845 m/s
h1 h2 h3 h4 5.0 5.0 10.0 10.0
     
v s ,1 v s , 2 v s , 3 v s , 4 375 646 1030 2126

though the site consists of a shallow soil layer that will lead to a completely different
response of the soil profile. In fact, the response of the soil profile will depend on
the largest impedance contrast between two layers, i.e. between Layer 2 and Layer 3.
A consideration of average shear wave velocity up to a depth to 10 m would be more
appropriate. The respective vs,10 would classify the site into soil site class B:
10 10
v s ,10    475 m/s
h1 h 5.0 5.0
 2 
v s ,1 v s , 2 375 646

12
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

1.2.3 Standard Penetration Test value NSPT


The Standard Penetration Test (SPT) can only be applied to soils that have a certain resistance
(applies to Ground types B – D). Due to SPT’s working principle, it cannot used for rock materials.
SPT is the oldest and most commonly used in-situ test procedure in geotechnical engineering.

Following Ohba and Toriumi (1970), the correlation between N and vs is:

v s  69  N 0.17  z 0.2  F in [m/s] eq. (1.5)

with: z – depth of material in [m]


F – factor between 1.0 – 1.45
(e.g. 1.1 for sand, 1.2 for gravel)

Further correlations between vs and N are provided by Islam (2005; Table 1.1).

Table 1.1. Empirical relationships between vs and N (Islam, 2005)).

Reference Equation Required parameters

Imai and Yoshimura (1970) v s  76  N 0.33


Ohba and Toriumi (1970) v s  84  N 0.31
F1 = 1.0 for Holocene soils
= 1.3 for Pleistocene soils
F2 = 1.00 for clay
= 1.09 for fine sand
Ohta and Goto (1978) v s  69  N 0.17  D 0.2  F1  F2
= 1.07 for medium sand
= 1.14 for coarse sand
= 1.15 for gravelly sand
= 1.45 for gravel
A = 102 b = 0.29 for Holocene clay
= 81 = 0.33 for Holocene sand
Imai (1977) vs  a  N b
= 114 = 0.29 for Pleistocene clay
= 97 = 0.32 for Pleistocene sand
Okamoto et al. (1989) v s  125  N 0.3 for Pleistocene sand
Tamura and Yamazaki
v s  105.8  N 0.187  D 0.179 D – depth in [m]
(2002)

13
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

1.2.4 Average Standard Penetration Test value NSPT


If shear wave velocities are not available the classification of the site shall be done based on NSPT
values (EN1998-1:2004, clause 3.1.2(2)). The calculation of average NSPT is not provided in
Eurocode 8. However, following provisions of U.S. building code IBC-2006 this is done in the
following way:
30
N SPT  N
in [blows/0.3 m] eq. (1.6)
hi

i 1 N SPT , i

with: hi – thickness of layer i in [m]


NSPT,i – Standard Penetration Test values of layer i in [blows/0.3 m]

1.2.5 Average undrained shear resistance


If the soil profile consists of cohesion material (clay and silt), the following equation can be used to
calculate the average undrained shear resistance of soil profile cu,30.
30
cu ,30  N
in [Kpa] eq. (1.7)
hi

i 1 cu ,i

Some recommendations:
- The direct shear strength should be used as undrained shear resistance, if the Norwegian ADP
method will be used.
- If the soil profile consists of both cohesion and cohesionless soil layers, it is suggested to use the
shear wave velocity to classify the ground. There are some correlations available between the
index soil properties of both cohesive and cohesionless soils (e.g. IP, cu, Dr, v0 etc.) to calculate
the shear wave velocity of soil.

14
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

1.2.6 Tutorial on site classification

Tutorial 1.4: Soil classification of model sites

Profile A Profile B Profile C

12
3.0 m clay
24
14
8 soft vs = 140 m/s
14.0 m clay 30
14 dense
16.0 m sand
6
36
15.0 m loose 6
sand 8
42
14
6
46
7
18
stiff N
7.0 m
dense clay cu = 120 kPa
> 7.0 very
m sand 26 > 16.0 m dense vs = 560 m/s
sand
40
rock
N > 10.0 m (sandstone)

Exercise: Which subsoil profile provides the most suitable conditions for the construction of a building with
shallow foundation solution ?
General notes and observations on each site:
- high liquefaction susceptibility - high impedance contrast at a - entire soil profile consists of
due to loose sands below water depth of 14.0 m will probably stiffer materials that will allow
table lead to larger soil amplification the use of a shallow foundation
- stiffer materials first available at a effects
depth of 18.0 m, which means - in order to reach stable soil,
that pile foundation would be piles or a deep basement would
necessary be necessary
- piling through liquefiable material
 Site C most suitable !
poses serious problems on pile
integrity/settlements

Identification of soil class:


- average SPT value < 15 - average vs,30 = 233 m/s - average SPT value of the upper
 Ground type C 16 m between 15–50
(NSPT ~ 13)  Ground type D
 Ground type C
- liquefiable soils are available
 Ground type S2 - undrained shear strength cu of
second layer between 70–250
 Ground type C

- NOTE: total thickness of rock-


overlying sediments is 23 m
if H  20 m  Ground type E!

15
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

1.3 Soil amplification


1.3.1 Effects of soil on ground motions (site effects)
Any soil profile acts as a filter modifying the amplitude, frequency content and duration of bedrock
motions. The general site effects can summarized as:
- the soil profile acts as filter,
- the amplitude of seismic waves increases so-called amplification (de-amplification of ground
motion can also happen in certain circumstances),
- the frequency content of bedrock motion changes and usually tends towards higher periods,
- the duration of ground motion changes, usually the soft soil profile tends to extend the duration
of ground motion.

Figure 1.2. Illustration of site effects.

1.3.2 Definition of soil amplification


The soil amplification factor S can be essentially defined in two ways:
1. Ratio of Fourier amplitudes:

a free  surface ( f )
S eq. (1.7)
a bedrock ( f )

2. Ratio of spectral amplitudes:

Sa , free  surface ( f )
S eq. (1.8)
Sa ,bedrock ( f )

1.3.3 Basin effect


It should be noted that ground motions are also strongly influenced by subsurface topography of
bedrock, though this is not considered explicitly in EN 1998. A major factor contributing to the
amplification of ground motion and increased damage in alluvium-filled valleys and basins is
interpreted to be the generation of surface waves at the valley edges and the reflection of these waves

16
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

back and forth through the soil layers infilling the valleys. This phenomenon is referred to as the
basin effect (Borcherdt and Glassmoyer, 1992; Finn et al., 1995).
Basin effects were interpreted to have significantly influenced the characteristics of the earthquake
ground motion and location of major damage centers in the Los Angeles and San Fernando Basins
during recent earthquakes (Borcherdt and Glassmoyer, 1992). The most common manifestations of
the basin effect are an increase in the duration and a shift to lower frequencies f (longer periods T)
that are more damaging for taller structures during an earthquake’s strong ground shaking.
With respect to the geotechnical and geological condition of Norway (shallow bedrock with sloping
face and variation of depth to bedrock) under many construction sites will probably the basin effects
play a significant rolls in amplifying the bedrock motions. Therefore a 2D site response analysis will
reasonably produce the more realistic amplification factor (S factor).

1.4 Site-specific amplification factor


If the ground type is defined as S1 or S2 it is required to perform a site-specific ground response
analysis to account for seismic action and design response spectrum at the site. The selection of the
procedure (equivalent linear versus nonlinear) and its complexity (1D versus 2D or 3D) is dictated by
the site and soil conditions. This chapter is meant informative and has a goal to give some guidance
for performing such ground response analysis. However, it is strongly recommended to involve an
earthquake geotechnical expert for doing such analysis.

1.4.1 Ground response analysis techniques


- Linear analyses (constant vs and D),
- Equivalent linear analyses:
a. Based on iterative procedure to establish strain-compatible dynamic soil properties. The
heart of the approach is the degradation curves, i.e. dynamic properties versus strain.
b. The pseudo-nonlinear approach removes a significant amount of high frequency energy
relative to the linear approach, see Figure 1.3.
c. Equivalent linear motions provide a more realistic approximation of motions (relative to
linear methods) for cases where materials do not behave elastically (i.e., soft rock and
soils), see Figure 1.4.
- Nonlinear analyses:
a. Nonlinear codes are typically required for cases where there is a high degree of strain in
the soil or significant pore pressure built up.
b. Choose a constitutive model representing nonlinear cyclic soil behavior (nonlinear
inelastic, cyclic plasticity, pore pressure generation).
c. Integrate the equation of motion for vertically propagating shear waves in time domain.

17
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

Figure 1.3. Equivalent linear procedure for ground response analysis (Rix, 2001).

Figure 1.4. Comparison linear and equivalent linear approaches.

1.4.2 Recommendations on performing site-specific ground response analysis


The ground response analysis is performed to:
- assess the site-specific amplification factor,
- estimate the seismic action at ground surface as well as within soil profile,
- provide primarily response spectra, for design and safety evaluation of structures,
- evaluate development of excess pore water pressure and potential for liquefaction,
- conduct first analytical phase of seismic stability evaluations for slopes and embankments,
- evaluate the seismic induced stress and strain within soil profile,
- estimate the parameters needed for foundation design and soil – structure interaction analysis.

Steps in site-specific ground response analysis


The procedure for performing the ground response analysis is summarized in Figure 1.5.

18
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

Figure 1.5. Site-specific ground response analysis procedure.

1. Characterization of the site:


The dynamic site characterization should be provided prior to run the analysis which includes the
following:
 Shear wave velocity profile
 Variation of shear modulus with strain (or modulus reduction curve).
 Variation of damping with strain (or damping ratio curve).

2. Selection of design input motions at bedrock:


Clause 3.2.3.1.2 and 3.2.3.2.3 of EN1998-1:2004 addresses the choosing of artificial and registered
ground motions. Appropriate rock motions (either natural or synthetic acceleration time histories) are
selected to represent the design rock motion for the site. The rock motion should be associated with
the specific seismotectonic structures, source areas or provinces that would cause most severe
vibratory ground motion or foundation dislocation capable of being produced at the site under
currently known tectonic framework.
If natural time histories are used, it is preferable to use a set of natural time histories that have
ground motion characteristics similar to those estimated for the design rock motions. That means the
selected histories should have:
 Peak ground motion parameters
 Response spectral content and
 Duration of strong shaking
In the absence of natural motions, artificial motions can be generated using the concept of ‘spectrum
compatible time histories’. For this problem several procedures are available such as time domain,
frequency domain generation, empirical Green’s function technique, etc.

3. Ground response analysis and design spectra


Ground response analysis (linear, equivalent linear or nonlinear) are performed for the site specific
profiles using the rock motions as input motion, to compute the time histories at the ground surface.

19
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

Response spectra of calculated ground surface motions are statically analyzed or interpreted in some
manner to develop ‘design spectrum’ for the site. The time histories from the ground response
analysis can be used directly to represent the ground surface motions or artificial time histories can
be developed to match the design spectrum.

4. Wave propagation analysis/site amplification


During earthquakes, the ground motion parameters such as amplitude of motion, frequency content
and duration of the ground motion change as the seismic waves propagate through overlying soil and
reach the ground surface. The phenomenon, wherein the local soils act as a filter and modify the
ground motion characteristics, is known as ‘soil amplification problem’. Physically, the problem is to
predict the characteristics of the seismic motions that can be expected at the surface (or at any depth)
of a soil stratum.

1.4.3 2D/3D site response analysis


A variety of finite element and finite difference computer programs are available for use in two-
dimensional seismic site response analyses. The computer program QUAD4 (Idriss et al., 1973), and
updated version QUAD4M (Hudson et al., 1994), is among the most commonly used computer
programs for 2D site response analyses.
Other 2D equivalent linear seismic site response analysis computer programs are available to public
including TELDYN (Pyke, 1995) and FLUSH (Lysmer et al., 1975).
A few modern, user friendly site response analyses programs are also available including QUAKE/W
and FLAC (2D/3D). Both programs have possibility to use the nonlinear soil model. However,
FLAC is more flexible with using the user defines constitutive soil models.

1.5 Topographic amplification


The topic on topographic amplification is briefly addressed in Chapter 4 and Informative Annex A
of EN 1998-5:2004. However, its should be mentioned that not much progress has been made on
this problem during the last years. Other building codes, e.g. AFPS (1990), are much more precise
than the provisions given in EC8.
In contrast to subsoil topography, more serious effects are reported to be caused by surface
topography. According to numerous reconnaissance studies after strong earthquakes, an increase of
damage to buildings can be observed on steep slope situations which extend towards the plateau.
According to different scientific groups (Aki, 1988; Bard, 1995, 1997; Géli et al., 1988) that deal with
instrumental and theoretical investigations of surface topography on ground motion characteristics,
the following can be stated (taken from Lang, 2004):
- Mountain tops or ridge crests, and more generally, convex topographies (such as cliff borders),
lead to an amplification of seismic ground motion, while valleys or foothills (concave
topographies) tend to deamplify the seismic signals.
- The effects of surface topography are larger on horizontal components than on vertical ones,
thus indicating that S motion is more affected by surface topography than P motion.

20
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

- The influence of surface topography on ground motion is directly related to the sharpness of
topography. Figure 1.6 illustrates this effect on the basis of two topographic features having
different wedge angles. According to this theoretical model, amplification of incoming seismic
waves increases as the wedge angle becomes sharper.
- Amplification and deamplification of seismic ground motion on topographic features are
apparently both frequency-dependent and band-limited: maximum effects can be observed for
wavelengths that approximately agree with the horizontal dimensions of the topographic shape.

Figure 1.6. Theoretical response of two topographical models having different wedge angles  to
vertically incident SH waves (after Bard, 1997).

EC8 suggests the use of a topographic amplification factor ST independent of period T, which shall
be used to increase the amplitude level of the derived acceleration spectrum. Topographic
amplification shall be considered, if:
(1) building of increased importance are concerned: γI > 1.0,
(2) ridge height H > 30 m,
(3) slope angle φ > 15º.

Two topographic features are addressed by EC8:


a) Isolated cliffs and slopes (Figure 1.7a)
b) Ridges with crest width significantly less than the base width (Figure 1.7b)
In case that a loose surface layer is present, the smaller value of ST shall be increased by at least 20%.
If a strongly irregular topography is present, a specific study is recommended.

ST ≥ 1.2 ST ≥ 1.2 ST ≥ 1.4

ST = 1.0 φ > 15º ST = 1.0 φ > 15º φ > 30º

(a) Islolated cliffs and slopes (b) Ridges

Figure 1.7. Topographic features and suggested amplification factors ST.

The topographic amplification factor ST is assumed to be 1.0 at the base and shall be linearly
interpolated between the top and the base.

21
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

1.6 Representation of seismic action


1.6.1 General
Earthquake action may be represented by a number of different ways. In Eurocode 8, both
horizontal and vertical seismic ground motion is represented by a smoothed acceleration response
spectrum. Its general shape and amplitude level are mainly dependent on the seismic hazard and the
local subsoil conditions of the respective site.

1.6.2 Horizontal elastic response spectrum vs. design spectrum for elastic analysis
EC8 distinguishes between the “horizontal elastic design spectrum” (chapter 3.2.2.2) and the
“design spectrum for elastic analysis” (chapter 3.2.2.5). Both are essentially representing the same
spectrum just with different parameters. While the elastic response spectrum is given in dependence
on damping correction factor , the design spectrum depends on behavior factor q. Since damping
correction factor  up- or downscales the spectral amplitudes according to the general work material
of the building under investigation (e.g. masonry, concrete, steel etc.), behavior factor q goes beyond
the material and also considers more specific structural features.
For  = 1.0 (i.e.  = 5%) and q = 1.0 both sets of equations (i.e. eq. (3.2)–(3.5) and eq. (3.13)–(3.16))
lead to exactly the same spectra except for the very short period range T < TB : The reason for this
lies in the fact that the design spectrum does not recognize overstrength in this period range and has
for behavior factors q ≥ 2.5 an unusual shape, which cannot be found in any other code. This,
however, only affects the spectral amplitudes for T < TB (see Figure 1.8).

Figure 1.8. Difference between the elastic response spectrum and the design spectrum for elastic analysis in
the short-period range.

The generation of both spectral types is straightforward and discussed in the Tutorial ‘Design
spectrum for elastic analysis’.
The spectrum is described by a number of parameters that depend on different impacts, such as:
1. Soil conditions: soil amplification factor S, corner periods TB, TC ,
2. Seismic hazard: ground acceleration on rock agR, corner period TD ,
3. Building occupancy: importance factor γI ,
4. Building typology: damping factor , damping correction factor , behavior factor q.
Spectrum-describing parameters are provided by the National Annex.

22
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

1.6.3 Tutorials on response spectrum generation

Tutorial 1.5: Horizontal elastic response spectrum (EN 1998-1:2004, 3.2.2.2)

Specifications: - building site Oslo (Norway)


- residential building  importance factor γI = 1.0
- ground type A

Determine design ground acceleration agR:


1) Identify ground acceleration value agR,40 Hz from Figure NA.3(901):
- the site of interest is located in a zone of agR,40 Hz > 0.5 m/s2 , but lower than 0.6 m/s2
 assume agR,40 Hz = 0.55 m/s2
2) Calculate design ground acceleration agR:
 agR = γI · 0.8 · agR,40 Hz = 1.0 · 0.8 · 0.55 = 0.44 m/s2

Derive spectral amplitudes Sa (T) for the horizontal design spectrum:


1) Soil parameters for Ground type A:
- based on Table NA 3.3: S = 1.0 , TB = 0.10 s, TC = 0.25 s, TD = 1.5 s
2) Building related parameters:
- elastic case  damping factor  = 5 % (damping correction factor  = 1.0)
3) Spectral amplitudes:
 T 
for 0 ≤ T ≤ TB : S a ( T )  a g  S  1   (  2.5  1)
 TB 
for TB ≤ T ≤ TC : S a ( T )  a g  S   2.5
T 
for TC ≤ T ≤ TD : S a ( T )  a g  S   2.5   C 
T 
T  T 
for TD ≤ T ≤ 4.0 s: S a ( T )  a g  S   2.5   C 2 D 
 T 

 calculation of Sa values for period range T = 0 – 3 sec (e.g. using MS Excel):

23
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

Tutorial 1.6: Vertical elastic response spectrum (EN 1998-1:2004, 3.2.2.3)

Specifications: - building site: north of Bergen (Norway)


- residential building  importance factor γI = 1.0

Determine horizontal design ground acceleration agR and calculate vertical design ground acceleration avg:
3) Identify ground acceleration value agR,40 Hz from Figure NA.3(901):
- the site of interest is located in a zone with agR,40 Hz > 0.9 m/s2
4) Calculate horizontal design ground acceleration agR :
 agR = γI · 0.8 · agR,40 Hz = 1.0 · 0.8 · 0.90 = 0.72 m/s2
5) Calculate vertical design ground acceleration avg :
 avg = 0.6 · agR = 0.6 · 0.72 = 0.432 m/s2

Derive spectral amplitudes Sve (T) for the vertical design spectrum:
4) Parameters for Norway:
- based on Table NA 3.4: TB = 0.05 s, TC = 0.20 s, TD = 1.2 s
5) Building related parameters:
- elastic case  damping factor  = 5 % (damping correction factor  = 1.0)
6) Spectral amplitudes:
 T 
for 0 ≤ T ≤ TB : Sve ( T )  a vg  1   (  3.0  1)
 TB 
for TB ≤ T ≤ TC : Sve ( T )  a vg   3.0
T 
for TC ≤ T ≤ TD : Sve ( T )  a vg   3.0   C 
T 
T  T 
for TD ≤ T ≤ 4.0 s: Sve ( T )  a vg   3.0   C 2 D 
 T 

 calculation of Sve values for period range T = 0 – 3 sec (e.g. using MS Excel):

24
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

Tutorial 1.7: Horizontal design spectrum for elastic analysis (EN 1998-1:2004, 3.2.2.5)

Specifications: - building site Oslo (Norway)


- residential building  importance factor γI = 1.0
- ground type A
- behavior factor q = 2.0
Determine design ground acceleration agR:
6) Identify ground acceleration value agR,40 Hz from Figure NA.3(901):
- the site of interest is located in a zone of agR,40 Hz > 0.5 m/s2 , but lower than 0.6 m/s2
 assume agR,40 Hz = 0.55 m/s2
7) Calculate design ground acceleration agR:
 agR = γI · 0.8 · agR,40 Hz = 1.0 · 0.8 · 0.55 = 0.44 m/s2

Derive spectral amplitudes Sd (T) for the horizontal design spectrum:


7) Soil parameters for Ground type A:
- based on Table NA 3.3: S = 1.0 , TB = 0.10 s, TC = 0.25 s, TD = 1.5 s
8) Spectral amplitudes:
 2 T 2.5 2 
for 0 ≤ T ≤ TB : Sd ( T )  a g  S     (  )
 3 TB q 3 
2.5
for TB ≤ T ≤ TC : Sd ( T )  a g  S 
q
2.5  TC 
for TC ≤ T ≤ TD : Sd ( T )  a g  S  
q  T 
   a g = 0.09 m/s2 with β = 0.20 (NA 3.2.2.5(4)P)
2.5  TC  TD 
for TD ≤ T ≤ 4.0 s: Sd ( T )  a g  S  
q  T 2 
   a g = 0.09 m/s2 with β = 0.20 (NA 3.2.2.5(4)P)

 calculation of Sd values for period range T = 0 – 3 sec (e.g. using MS Excel):

25
EC8 Guide Book Chapter 1 – Ground conditions and seismic action

26
EC8 Guide Book Chapter 2 – Seismic design principles

2 Seismic design principles


Dominik H. Lang and Emrah Erduran (NORSAR, Kjeller)

2.1 Basic principles of conceptual design


2.1.1 Structural simplicity

Principle: Clear and direct load paths so that seismic forces can be transferred
from top to bottom.

This directly implies that the structure should be regular both in plan and elevation. All elements of
the load path should have adequate strength and ductility.

Figure 2.1. Direct load paths from top to bottom can only be assured by
regularly arranged structural elements.

2.1.2 Uniformity, symmetry and redundancy

Principle: The greater the departure from regularity or symmetry, the greater
the peril to suffer serious damage!

During the design process care should be taken with respect to the following:
→ symmetrical arrangement of mass and stiffness in ground plan and elevation
→ irregularities in the buildings layout will result in eccentricities of mass and stiffness center, which
will lead to unfavorable torsional effects that are coupled with large displacements
→ interruptions of lateral stiffness over the buildings height will cause weak locations in the
structural system

27
EC8 Guide Book Chapter 2 – Seismic design principles

Redundancy means that the building should have more than one single load path in order to
transfer the forces from top to bottom. In case that one load path is affected by the shaking, another
load path is taking over. Redundant systems are generally less simple.

(a) Regularity in plan


Figure 2.2 opposes the most common regular with irregular plan shapes. However, a regular plan
shape cannot always be realized. Consequently, measures have to be taken in order to provide that an
irregular plan shape ‘behaves’ like a regular one. This can be either done by:
- subdividing the building into separate independent units using seismic joints (Figure 2.3),
- strengthening those parts of the building that experience largest displacements/forces (Figure 2.3).

(a) irregular (b) regular


Figure 2.2. Examples for irregular and regular plan shapes.

(a) irregular (b) irregular with improvements


Figure 2.3. Improving irregular plan shapes.

Regular plan shapes with regularly arranged structural elements are the prerequisite for torsional
resistance (Chapter 2.1.4). Torsional effects predominantly occur if the center of mass (M) and the
center of stiffness (S) do not fall into the same location and thus exhibit an eccentricity (e; Figure
2.4). This may be either produced by:
- irregular plan shapes
- regular plan shapes with irregularly distributed structural elements such as columns, walls or cores
(Figure 2.5).

28
EC8 Guide Book Chapter 2 – Seismic design principles

(a) M  S → torsional motion (b) M = S → pure lateral motion


Figure 2.4. Torsional movements are to be avoided as they lead to larger relative displacements
as compared with pure lateral motions.

(b) avoiding torsional motion through structural


(a) M  S → torsional motion
arrangement
Figure 2.5. Torsional movements may also occur in buildings that show plan regularity at first sight.

b) Regularity in elevation
Figure 2.6 compares favorable with unfavorable building solutions with respect to elevation shape.
In general the following principles shall be considered:
- avoid sensitive zones where concentrations of stress or large ductility demands increase damage
susceptibility,
- provide either constant or continuously decreasing stiffness with height,
- avoid buildings with too slender (chimney-style) or inverted pendulum shape (heavy mass on top),
- arrange continuous bracings over total building height,

29
EC8 Guide Book Chapter 2 – Seismic design principles

- arrange seismic (movement) joints of sufficient thickness1 in order to decouple separate building
segments from each other.

Damage susceptibility increases dramatically if the building’s lateral stiffness is interrupted over
height, e.g. by the arrangements of soft stories (‘soft story effect’; Figure 2.7). Further, the spatial
dislocation of the stiffness axis may be affected if e.g. strengthening measures are erroneously applied
(Figure 2.8). Care should also be taken with respect to the construction of buildings on sloped
terrain which may lead to different column wall heights, or with respect to equal floor masses over
height (Figure 2.8). Heavy masses on top, e.g. by the arrangement of a water tank shall always be
avoided.

(a) unfavorable elevation solutions (b) favorable elevation solutions

Figure 2.6. Principles of regularity in elevation.

(a) irregular elevation shape


(b) regular elevation shape
through arrangement of soft stories
Figure 2.7. Interruptions of lateral stiffness over height by ‘soft stories’.

1 the thickness of seismic joints shall be designed so that pounding effects between both building are avoided

30
EC8 Guide Book Chapter 2 – Seismic design principles

(a) irregularities in elevation (b) regular elevation shapes


Figure 2.8. Interruptions of lateral stiffness over height by strengthening of certain building parts,
irregular foundation heights or irregular story masses.

2.1.3 Bi-directional resistance and stiffness


Earthquake ground motion is a spatial phenomenon and buildings which are designed to withstand
these loads shall possess bi-directional resistance and stiffness. It should always be carefully checked
whether RC beams and/or bracing walls are available in both general building directions. Often
supporting walls are left or removed because of the arrangement of large openings, e.g. for shops.

A special case consists in slender plan shapes: Please consider that both general building axes usually
possess equal resistance. A ‘weak’ axis (as sometimes reported) does not exist (compare with Figure
2.9).

H1

H2

Figure 2.9. Slender plan shape. Both general building axes possess equal resistance.

2.1.4 Torsional resistance and stiffness


Torsional effects generally occur if the center of mass and the center of stiffness are located in a
certain distance (i.e. eccentricity e) to each other. This is caused by:

31
EC8 Guide Book Chapter 2 – Seismic design principles

- symmetrical/regular plan shapes where the structural elements are irregularly distributed (Figure
2.10),
- unsymmetrical/irregular plan shapes where masses and stiffnesses are disorderly arranged.

center of stiffness S eccentricity e

center of mass M

Figure 2.10. Regular plan shapes but irregularly distributed structural elements.

It should be also considered that slender plan shapes tend more easily to torsional effects (Figure
2.11). It is therefore advisable to limit the length-to-width ratio (slenderness) of the plan:
 = Lmax/Lmin < 4.0 eq. (2.1)
With regard to other seismic provisions the threshold value given by EC8 is estimated as too high
and should probably limited to 2.5.
In general, in order to avoid the occurrence of torsional effects, lateral resisting structural elements
shall be distributed close to the periphery of the building. Further, the provisions given with respect
to REGULARITY IN PLAN are important.

Lmin

Lmax

Figure 2.11. Slender plan shape tend more easily to torsional movements.

2.1.5 Diaphragmatic behavior at storey level


The main role of floor diaphragms is to collect and transmit the inertia forces. They further have to
ensure that all vertical elements act together (are synchronous) in resisting the seismic forces and

32
EC8 Guide Book Chapter 2 – Seismic design principles

should thus have sufficient in-plane stiffness. In order to ensure this, large openings or interruptions
should be avoided (Figure 2.12).
Floor diaphragms act as horizontal ties, which prevent excessive relative deformations between the
vertical elements.

(a) unfavorable arrangement of diaphragms (b) favorable solutions

Figure 2.12. Regularity of floor slabs in order to ensure diaphragm action.

2.1.6 Adequate foundation


The foundation shall ensure that the whole building is subjected to a uniform seismic excitation. If
the superstructure is likely to differ in width and stiffness, a rigid, box-type or cellular foundation
should be chosen (Figure 2.13).
The provisions of EN 1998-5:2004 have to be taken into account. Further, the chapters on
geotechnical issues and soil-structure interaction have to be considered.

(a) unfavorable foundation solutions (b) favorable foundation solutions

Figure 2.13 Regularity in foundation.

33
EC8 Guide Book Chapter 2 – Seismic design principles

2.2 Primary and secondary seismic members


While primary elements contribute to the seismic resistance, secondary elements resist gravity loads
only and have no contribution to the structure’s seismic resistance (Figure 2.14).
According to EC8, the contribution of all secondary seismic members should not exceed 15 % of
that of all primary seismic members in order to avoid 'unintentional stiffening' by secondary elements
(higher frequencies and inertial loads) as well as to provide flexible joints.

primary members

secondary members

Figure 2.14 Regular plan shapes but irregularly distributed structural elements.

2.3 Criteria for structural regularity


2.3.1 Plan regularity
Repeating the principles given in Chapter 2.1.2 a), criteria for plan regularity can be summarized as
follows:
- symmetrical arrangement of lateral stiffness and mass distributions,
- two orthogonal axes, i.e. rectangular buildings with two principal building axes that are
orthogonally arranged to each other,
- plan configuration is compact, able to be delimited by a polygonal convex line (compare with
Figure 2.15),
- if set-backs exist, plan can be estimated as 'compact' if the differential area is less than 5 % of the
total floor area (compare with Figure 2.16 ; i.e. A1 < 0.05 · Atot and A2 < 0.05 · Atot),
- slenderness , i.e. ratio between length Lmax and width Lmin of the building shall be smaller than 4.02
(Figure 2.17).

2 the threshold for slenderness  > 4.0 is estimated too high and should better be set to 2.5! Please consider also that
buildings of slender, long drawn-out plan shapes tend to be affected by torsional effects more easily.

34
EC8 Guide Book Chapter 2 – Seismic design principles

(b) pure polygonal convex line and the most regular


(a) polygonal convex and concave line
convex plan shapes
Figure 2.15. Difference between polygonal concave and convex lines.

A1

convex perimeter line plan shape (Atot) is compact, if:


Atot of the building A1 < 5.0 % of Atot
and A2 < 5.0 % of Atot

A2

Figure 2.16. Criteria for ‘compactness’ of plan shapes with setbacks.

plan shape is slender, if:


Lmin
 = Lmax / Lmax > 4.0
Lmax

Figure 2.17. Criteria for ‘compactness’ of plan shapes with setbacks.

Plan shapes are further considered regular, if the structural eccentricity eo for each story level and in
both principal building axes x and y is smaller than a certain ratio of the torsional radius rx and ry,
respectively:
eox  0.30 · rx eq. (2.2)
while rx  ls !!
with:
eox  distance between the centre of stiffness and the centre of mass (in each direction)
rx  square root of the ratio of the torsional stiffness to the lateral stiffness in the y direction (“torsional
radius”)
ls  is the radius of gyration of the floor mass in plan (square root of the ratio of (a) the polar moment of
inertia of the floor mass in plan with respect to the centre of mass of the floor to (b) the floor mass)

35
EC8 Guide Book Chapter 2 – Seismic design principles

y
coordinates: x, y EIx

EIy

xcS S

ycS

Figure 2.18. Criteria for ‘compactness’ of plan shapes with setbacks.

To check the criteria of eq. (2.2), a number of parameters have to be calculated. This is successively
done in the following:
(1) Center of stiffness S (coordinates xcS and ycS):
( x  EI x )
x cS   eq. (2.3)
EI x

( y  EI y )
y cS   eq. (2.4)
EI y

(2) Eccentricity eo between S and M:


eox = |xcS  xcM| in x direction eq. (2.5)

eoy = |ycS  ycM| in y direction eq. (2.6)


(3) Torsional radius rx :

( x 2  EI x  y 2  EI y )
rx   in x direction eq. (2.7)
 ( EI x )

( x 2  EI x  y 2  EI y )
ry   in y direction eq. (2.8)
 ( EI y )

(4) Radius of gyration ls :

(a 2  b 2 )
ls  for rectangular plan shapes eq. (2.9)
12

36
EC8 Guide Book Chapter 2 – Seismic design principles

The calculation of the radius of gyration for irregular plan shapes:

y A2
xM,2 M2
b2

A1 a2

xS
S yM,2
xM,1 M1
b1

yM,1 yS

x
a1

For subarea A1 :
l 12  ( x M ,1  x M )2  ( y M ,1  y M )2 in [m2] eq. (2.10)
For subarea A2 :
l 22  ( x M , 2  x M )2  ( y M , 2  y M )2 in [m2] eq. (2.11)
Radius of gyration ls :

( a1b13  a13b1 ) /12  a1b1l 12  ( a 2b 23  a 23b 2 ) /12  a 2b 2l 22


ls  in [m] eq. (2.12)
a1b1  a 2b 2

37
EC8 Guide Book Chapter 2 – Seismic design principles

2.3.2 Elevation regularity

Principle: “Both the lateral stiffness and the mass of the individual stories shall
remain constant or reduce gradually, without abrupt changes, from
the base to the top of a particular building.” (4.2.3.3(3))

Elevation regularity is in principal defined through the dimension of setbacks, if present. The
conditions described in 4.2.3.3 (5) and restructured by Table 2.1 apply.

Table 2.1 Criteria for regularity in elevation with respect to the arrangement of setbacks.

Case Illustration Conditions

1 – gradually and - the setback at any floor shall not be greater than 20% of
symmetrical arranged the plan dimension below:
setbacks
L1  L 2
criterion:  0.20
L1
2 – setback above 0.15·H - if setback occurs above a height that corresponds to more
than 15 % of the total height, the sum of both setbacks
shall not be greater than 20% of the base plan dimension:
L 3  L1
criterion:  0.20
L

3 – setback below 0.15·H - if setback occurs at a height that is within (below) the 15
% of the total height, the sum of both setbacks shall not
be greater than 50% of the base plan dimension:
L 3  L1
criterion:  0.50
L
- the structure of the base zone within the vertical
projection of the upper stories (i.e. L–(L3+ L1)) shall be
designed to resist 75% of the shear forces that would
develop in a similar building without the base
enlargement (i.e. L–(L3+ L1))
4 – asymmetrical arranged - the sum the setbacks at all stories shall not be greater than
setbacks 30% of the base plan dimension:
L  L2
criterion (1):  0.30
L
- the individual setbacks shall not be greater than 10% of
the plan dimension below:
L1  L 2
criterion (2):  0.10
L1

38
EC8 Guide Book Chapter 2 – Seismic design principles

2.3.3 Tutorial on plan regularity

Tutorial 2.1: Plan regularity of an irregular shear wall system

Plan layout:
6.0 m

y
6.3  0.3 m2
6.0 m

31.74 m
S
19.80 m
M

0.3  12.3 m2
9.00 m
7.80 m
 0.3  6.3 m2

x
6.3  0.3 m2

Calculation of the stiffness center S:


1. Moments of inertia I for the single shear walls:

h b3 4 b h3 4
Wall no. I ix  [m ] xi [m] I iy  [m ] yi [m]
12 12
 6.25 0.0 0.0 3.0
 0.0 27.0 6.25 0.0
 0.0 21.0 6.25 18.0
 46.52 36.0 0.0 6.0
 Iij 52.77 – 12.50 –

2. Coordinates of the stiffness center S:

( x  EI x ) ( 6.25  0.0 )  ( 46.52  36.0 )


x cS     31.74 m
EI x 52.77

( y  EI y ) ( 6.25  0.0 )  ( 6.25 18.0 )


y cS     9.00 m
EI y 12.50

 Location of stiffness center S is indicated in sketch above.

39
EC8 Guide Book Chapter 2 – Seismic design principles

Tutorial 2.1: Plan regularity of an irregular shear wall system

Calculation of the mass center M:


3. Coordinates of the mass center M:
( x A1  A1 )  ( x A2  A2 ) ( 9.0  (18.0 12.0 ))  ( 27.0  (18.0 18.0 ))
x cM    19.80 m
Atotal ((18.0 12.0 )  (18.0 18.0 ))
( y A1  A1 )  ( y A2  A2 ) ( 6.0  (18.0  12.0 ))  ( 9.0  (18.0  18.0 ))
y cM    7.80 m
Atotal ((18.0  12.0 )  (18.0  18.0 ))
Check for plan regularity:
4. Eccentricity eo between S and M:
- in x direction: eox =|xcS  xcM|= 31.74 – 19.80 = 11.94 m
- in y direction: eoy =|ycS  ycM|= 9.00 – 7.80 = 1.20 m

5. Calculation of torsional radius r :


- in x direction:
( x 2  EI x  y 2  EI y ) ( 31.74 2  6.25)  (4.26 2  46.52 )  ( 9 2  6.25)  ( 9 2  6.25)
rx      12.43 m
 ( EI x ) 52.77

- in y direction:
( x 2  EI x  y 2  EI y ) ( 31.742  6.25)  (4.26 2 46.52 ) ( 92  6.25)  ( 92  6.25)
ry      25.54 m
 ( EI y ) 12.50

6. Radius of gyration ls :
l 12  ( x 1  x cm ) 2  ( y1  y cm )2  (18  19.8)2  ( 6  7.8 )2  6.48m 2
l 12  ( x 2  x cm ) 2  ( y 2  y cm )2  ( 27  19.8 ) 2  (15  7.8 )2  103.68m 2
( a1b13  a13b1 ) /12  a1b1l 12  ( a 2b 23  a 23b 2 ) /12  a 2b 2 l 22
ls  
a1b1  a 2b 2
( 36 12 3  36 3 12 ) /12  36 12 6.48  (18 6 3  183  6 )/12 18 6 103.68
  11.31m
36 12  18 6

7. Criteria for plan regularity:


- in x direction: eox  0.30 · rx  11.94 m  0.30 · 12.43 m = 3.73 m 
if rx  ls :  rx = 11.94 m  11.31 m 
- in y direction: eoy  0.30 · ry  1.20 m  0.30 · 25.54 m = 7.66 m 
if ry  ls :  ry = 25.54 m  11.31 m 

 As the condition is not fulfilled for both principal building axes, the building is to be
estimated as irregular in plan!

40
EC8 Guide Book Chapter 3 – Soil–structure interaction

3 Soil-structure interaction
Farzin Shahrokhi (Multiconsult, Oslo)

3.1 Introduction
EN1998-5:2004 indicates that the effects of dynamic soil-structure interaction shall be taken into
account in some cases (ref. NS1998-5, section 6), but it does not specify details how SSI is analyzed.
Two types of SSI are commonly referred to in the literature:
1. “Kinematic” interaction is caused by inability of a foundation to follow ground motion due
to greater foundation stiffness in comparison with ground stiffness. In effect, stiff foundation
filters high frequency ground motion to an averaged translational and rotational foundation
motion. Average values are smaller than the maximum values and therefore “kinematic”
interaction is beneficial except if averaged motion results in significant rotation and rocking
of a foundation.
2. “Inertial” interaction is caused by the existence of structural and foundation masses. Seismic
energy transferred into a structure is dissipated by material damping and radiation back into
ground causing superposition of incoming and outgoing ground waves. As a result, the
ground motion around a foundation can be attenuated or amplified, depending on a variety
of factors.
The most important factor in determining the response is the ratio between the fundamental period
of a foundation and the fundamental period of the adjacent ground in the free-field. The ratio of
unity indicates resonance condition between foundation and its adjacent ground, which is to be
avoided.

3.2 Code approach to Soil Structure Interaction analyses


In almost every seismic building code, the structural response and foundation loads are computed
neglecting the soil-structure interaction that is a fixed base analysis of the structure is performed. The
belief is that SSI always plays a favorable role in decreasing the inertia forces; this is clearly related to
the standard shape of code design spectra, which almost invariably possesses a gently descending
branch beyond a constant spectral acceleration plateau as presented in Figure 3.1.
Lengthening of the period, due to SSI, moves the response to a region of smaller spectral
accelerations. However there is evidence that some structures founded on unusual soils are
vulnerable to SSI.

41
EC8 Guide Book Chapter 3 – Soil–structure interaction

2.0

Sd
1.5 Rigid base
Flexible base
1.0
Sd (m/s2)

0.5

0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

T (s)

Figure 3.1. Illustration of the effect of SSI on period lengthening and seismic demand.

Section 6 of EN1998-1:2004 states that "the effects of dynamic soil-structure interaction shall be
taken into account in the case of:
• structures where P– effects play a significant role;
• structures with massive or deep seated foundations;
• slender tall structures;
• structures supported on very soft soils, with average shear wave velocity less than 100 m/s.
• The effects of soil-structure interaction on piles shall be assessed."

In addition, Annex D of EN1998-5:2004 describes the general effect of SSI and a specific chapter
analyzes its effects on piles and the way to deal with it.
EN1998-1:2004 is the only code, which recognizes the importance of kinematic interaction for piled
foundations, as it is stated in clause 5.4.2(6) of EN1998-5:2004:
"Bending moments developing due to kinematic interaction shall be computed only when two or
more of the following conditions occur simultaneously:
• the subsoil profile is of class C (soft soil), or worse, and contains consecutive layers with
sharply differing stiffness,
• the zone is of moderate or high seismicity, S  ag > 0.1 g,
• the supported structure is of importance category III or IV."

3.3 Basic concepts of SSI


Unfortunately, the approach traditionally used in engineering practice to account for the SSI problem
is often oversimplified and is concerned mostly by modeling the static deformability of the ground.
It should be emphasized that the interest in studying the seismic SSI is motivated not only by the
need to satisfy geotechnical requirement related to foundation response to earthquake loading (e.g.
bearing capacity assessment, settlement calculation), but also by the necessity of computing the
“effective” earthquake excitation to a structure with respect to the free-field ground motion (which is
also called Foundation Input Motion or FIM).

42
EC8 Guide Book Chapter 3 – Soil–structure interaction

The latter strongly influences the structure seismic demand as it takes into account both the soil-
foundation coupling (i.e. the dynamic impedance function of the soil-foundation system) and
scattering effects caused by the motion of embedded foundation (i.e. the kinematic interaction
effects).
It is rather unfortunate that kinematic interaction effects are often neglected in engineering practice
due to the difficulties of their evaluation even though they may be relevant.

There are three primary categories of soil-structure interaction (SSI) effects. These include:
• filtering of the ground motions transmitted to the structure (kinematic effects),
• introduction of flexibility to the soil-foundation system (flexible foundation effects),
• dissipation of energy from the soil-structure system through radiation and hysteretic soil
damping (foundation damping effects).

In the SSI literature, foundation stiffness and damping effects are often described in terms of
impedance functions. The impedance function should account for the soil stratigraphy and foundation
stiffness and geometry, and is typically computed using equivalent-linear soil properties appropriate
for the in situ dynamic shear strains, clause 4.2.3(3) of EN1998-5:2004.
Impedance functions can be evaluated for multiple independent foundation elements or (more
commonly) a single 6×6 matrix of impedance functions is used to represent the complete foundation
(which assumes foundation rigidity).

Figure 3.2a illustrates the assumption that the structural model is founded on a rigid base that is
excited by the free-field motion. The fixed-base modeling assumption is inappropriate for many
structures. Structural systems that incorporate stiff vertical elements for lateral resistance (e.g., shear
walls, braced frames) can be particularly sensitive to even small base rotations and translations that
are neglected with a fixed base assumption.

Figure 3.2b illustrates the incorporation of foundation flexibility (stiffness) into the structural model
directly. Most provisions use the free-field motion as the design seismic demand with 5% damping as
the conventional initial value.
This approach is capable of modeling both the structural and geotechnical (soil) components of the
foundation. The result is that the response of the overall structural system includes deformations in
the structural and geotechnical parts of the foundation system.
Compared with the fixed-base modeling approach:
• the predicted period of the structure lengthens,
• the distribution of forces among various elements changes,
• the sequence of inelasticity and the modes of inelastic behavior can change,
• foundation mechanisms (e.g., rocking, soil bearing failure, and pier/pile slip) can be directly
evaluated and considered.
All of these effects result in more realistic evaluation of the probable structural behavior and
performance.

43
EC8 Guide Book Chapter 3 – Soil–structure interaction

Figure 3.2c illustrates the filtering effects that soil-structure interaction can have on the character
and intensity of ground motion experienced by the structure. Kinematic interaction results from the
presence of relatively stiff foundation elements on or in soil that cause foundation motions to deviate
from free-field motions.
Two effects can be identified:
1. Base-slab averaging: The base-slab averaging effect can be visualized by recognizing that
motion that would have occurred in the absence of the structure is not the same at every
point. Placement of a structure and foundation across these spatially variable motions
produces an averaging effect in which the overall motion is less than the localized maxima
that would have occurred in the free field.
2. Embedment effects: The embedment effect is associated with the reduction of ground
motion that tends to occur with depth in a soil deposit.

Both base-slab averaging and embedment affect the character of the foundation-level motion
(foundation input motion, or FIM) in a manner that is independent of the superstructure (i.e., the
portion of the structure above the foundation), with one exception.
The effects are strongly period-dependent, being maximized at small periods. The effects can be
visualized as a filter of the free-field ground motion. The impact of those effects on superstructure
response will tend to be greatest for short-period buildings.
Figure 3.2d illustrates foundation damping effects that are another result of inertial soil-structure
interaction in addition to foundation flexibility. Foundation damping results from the relative
movements of the foundation and the supporting soil. It is associated with radiation of energy away
from the foundation and hysteretic damping within the soil. The result is an effective decrease in the
spectral ordinates of ground motion experienced by the structure.

44
EC8 Guide Book Chapter 3 – Soil–structure interaction

Figure 3.2. Foundation modeling assumptions (FEMA-440, 2005).

The foundation damping is linked to the ratio of the fundamental period of the system on the
flexible-foundation to that of a fixed-base model. Other factors affecting foundation damping are the
foundation size and embedment. The foundation damping is combined with the conventional initial
structural damping to generate a revised damping ratio for the entire system, including the structure,
foundation, and soil. This system damping ratio then modifies the foundation input motion imparted
to the system model as seismic shaking demand.

3.4 Step procedure to include SSI


An implemented step procedure to solve SFSI problem can summarized as given in Table 3.1.

Table 3.1. Procedure to solve SFSI problem.


1 Earthquake Scenario Consistent definition of rock-outcrop
2 Site Response Analysis 1D/2D equivalent linear/non-linear model, strain compatible dynamic soil
properties
3 Kinematic Interaction Foundation Input Motion (FIM), Foundation kinematic actions demand

4 Impedance Analysis Foundation stiffness and damping matrix

5 Inertia Interaction Dynamic transient structural analysis

6 Foundation feedback Foundation total seismic forces demand, soil and foundation capacity

45
EC8 Guide Book Chapter 3 – Soil–structure interaction

Step 1 and 2 are discussed in earlier chapters in this manual. Step 3, 4 and 6 belong to geotechnical
analysis, while step 5 is a dynamic structural analysis. Therefore, it is obvious that dynamic analysis of
structures including the SSI effects require a close cooperation and clear communication between the
structural and geotechnical engineers.

3.5 Kinematic effects


Kinematic interaction results from the presence of stiff foundation elements on or in soil, which
causes foundation motions to deviate from free-field motions as a result of base slab averaging and
embedment effects as illustrated in Figure 3.3.
Placement of a foundation slab across these variable motions produces an averaging effect in which
the foundation motion is less than the localized maximum that would have occurred in the free-field.
The embedment effect is simply associated with the reduction of ground motion that tends to occur
with depth in a soil deposit.
Theoretical models for kinematic interaction effects are expressed as frequency dependent ratios of
the Fourier amplitudes (i.e., transfer functions) of foundation input motion (FIM) to free-field
motion.
The FIM is the theoretical motion of the base slab if the foundation and structure had no mass, and
is a more appropriate motion for structural response analysis than the free-field motion.
Kinematic interaction effects tend to be important for buildings with relatively short fundamental
periods, large plan dimensions, or basements embedded 3 m or more in soil materials.
It is obvious in Figure 3.4 that the kinematic effects are more important for soft soils (ground type
D and S) rather than stiff soils.
A Ratio of Response Spectra (RRS) factor can be used to represent kinematic interaction effects. An
RRS is simply the ratio of the response spectral ordinates imposed on the foundation (i.e., the
foundation input motion, FIM) to the free-field spectral ordinates. Base slab averaging occurs to
some extent in virtually all structures.

Figure 3.3. Base slab averaging and embedment.

In following cases the kinematic effects should not been considered:


• In case of structures without a laterally connected foundation system and with flexible floor
and roof diaphragms the base-slab averaging effects should be neglected.

46
EC8 Guide Book Chapter 3 – Soil–structure interaction

• Foundation embedment effects should not be considered for structures without basements,
even if the footings are embedded.

Figure 3.4. Example of RRS for foundation embedded in different site class, b) with different embedment
depth (FEMA-440, 2005).

3.6 Foundation Damping


The damping from foundation-soil interaction is associated with hysteretic behavior of soil as
illustrated in Figure 3.6, as well as radiation of energy into the soil from the foundation (i.e.,
radiation damping) as illustrated in Figure 3.5. These foundation damping effects tend to be
important for stiff structural systems (e.g., shear walls, braced frames), particularly when the
foundation soil is relatively soft (i.e., site classes D and S1).
The effects of foundation damping are represented by a modified system-damping ratio. The initial
damping ratio for the structure neglecting foundation damping is referred to as i , and is generally
taken as 5%. The damping attributed to foundation-soil interaction alone (i.e., the foundation
damping) is referred to as f .
Finally, the damping ratio of the complete structural system, accounting for foundation-soil
interaction, as well as structural damping, is referred to as 0. The change in damping ratio from i to
0 modifies the elastic response spectrum. The spectral ordinates are reduced if 0 > i .
The following simplified equation can be used to estimate f and the subsequent spectral ordinate
change due to the modified damping ratio of the complete structural system, 0.

47
EC8 Guide Book Chapter 3 – Soil–structure interaction

b) c)
a)
Figure 3.5. Illustration of radiation damping.

a) b)
Figure 3.6. Illustration of hysteretic damping in soil in vicinity of foundation.

Figure 3.7. Example of foundation damping as a function of effective period lengthening ratio
(FEMA-440, 2005).

3.7 Analysis of impedance functions


Simplified impedance function solutions are available for shallow rigid circular and rectangular
foundations underlain by a uniform, viscoelastic half space.
Springs are used to present soil-foundation interaction. The foundation impedance function has a
complex form value. Combination of real and imaginary parts comprises the “Impedance function”:

48
EC8 Guide Book Chapter 3 – Soil–structure interaction

where: (dimensionless frequency)


 is Poisson’s ratio
j denotes both the translation (x or y direction) and rotation .

The real part represents stiffness and the imaginary part is related to damping. If foundation can be
considered as rigid, for 2D system, there are 3 springs support (horizontal, vertical and rocking) and
in 3D system, there should be defined 6 spring supports, see Figure 3.8.

Figure 3.8. Spring support in 2D model.

Introducing the stiffness coefficients affects the total response of the structure – foundation system
to seismic loading, see Figure 3.9. One of the most obvious effect is period lengthening for a flexible
base system compared to rigid base system. As discussed earlier the period lengthening is also related
to total system damping, as presented by following equations.

where: represent the foundation-structure stiffness,


and are translational and rotational stiffnesses of soil-foundation system.

Figure 3.9. Flexible base system.

49
EC8 Guide Book Chapter 3 – Soil–structure interaction

In the very basic form, the impedance functions can defines as static stiffness (e.g. ) and then
applying the dynamic modifiers (e.g. , ) to produce the dynamic stiffness as follows:

 Dynamic translational stiffness


 Dynamic rotational stiffness
 Damping coefficient for translational component

 Damping coefficient for rotational component

where: , , and are equivalent foundation radii,


is foundation area, and
is moment of inertia of footing.

3.7.1 Static stiffness of surface foundations


Static stiffness for a circular foundation on homogenous half space can approximate by:

Static stiffness of rectangular surface foundation on homogenous half space can approximate as
presented in Table 3.2:

Table 3.2. Static spring constant for rigid rectangular surface foundation on homogeneous half space
(FEMA-356)

50
EC8 Guide Book Chapter 3 – Soil–structure interaction

3.7.2 Embedment modification of static stiffness


For circular foundation at depth, e:

For rectangular foundation at depth, D, the equations are summarized in Table 3.3.

51
EC8 Guide Book Chapter 3 – Soil–structure interaction

Table 3.3. Foundation spring constant for rigid rectangular embedded foundation (FEMA-356).

3.7.3 Dynamic modification

Figure 3.10. Foundation stiffness and damping factors for elastic half space (Veletsos and Vebric, 1973).

The presented approach for shallow foundation and the equations presented in Annex C of EN1998-
5 for pile foundations are simplified methods and includes great limitations in practical use. There are

52
EC8 Guide Book Chapter 3 – Soil–structure interaction

some powerful FEM programs developed to calculate the frequency dependent foundation
impedance functions for both shallow and deep foundations which can account for soil layering.

3.8 Evaluation of seismic loads


3.8.1 Improved evaluation of seismic loads
Linear Systems
Referring to the multistep approach described previously, the last step of an SSI analysis (response of
the structure connected to the impedances) can be performed on a routine basis provided that:
• the system remains linear;
• the kinematic interaction can be neglected;
• dynamic impedance functions are readily available.
Although the superposition theorem is exact for linear soil, pile and structure, it can nevertheless be
applied to moderately nonlinear systems. This can be achieved by choosing reduced soil
characteristics which are compatible with the free-field strains induced by the propagating seismic
waves. This is the basis of the equivalent linear method, pioneered by Idriss and Seed (1968).
Table 4.1 in EN1998-5:2004 provides a transition between the linear elastic approach and nonlinear
methods. The mentioned table manifests that with increasing ground acceleration the soil adjacent to
a shallow foundation will experience increasing shear strains and consequently the stiffness will
decrease and the material damping increase, and it suggests how the apparent average shear modulus
and material damping of the soil adjacent will change with increasing peak ground acceleration and
envisages that an elastic SSI calculation would be done with the modified values for the soil stiffness
and damping.
It should be aware of following this simplification there is, of course, no frequency dependence on
the stiffness and damping parameters for the foundation.

Nonlinear Systems
One of the main limitations of the multistep approach is the assumption of linearity of the system for
the superposition theorem to be valid. As noted previously, some nonlinearity, such as those related
to the propagation of the seismic waves, can be introduced but the nonlinearities specifically arising
from soil-structure interaction are ignored.
The generic term "nonlinearities" covers geometrical nonlinearities, such as foundation uplift, and
material nonlinearities, such as soil yielding around the edges of shallow foundations, along the shafts
of piles, and the formation of gaps adjacent to pile shafts.
Those nonlinearities may be beneficial and tend to reduce the forces transmitted by the foundation
to the soil and therefore decrease the seismic demand. This has long been recognized for foundation
uplift for instance.

53
EC8 Guide Book Chapter 3 – Soil–structure interaction

3.9 Stiffness and capacity of pile foundation


The ultimate capacity of the piles, or pile groups, has to be checked once the applied inertia loads
acting at the pile cap are known. Two failure modes must be examined:
1. bearing failure with a vertical force exceeding the available tip and shaft resistance
2. lateral failure when the available lateral resistance is mobilized along part the pile shaft.
As well as capacity, the stiffness of the pile, or pile group, needs to be evaluated for the calculation of
kinematic and inertial interaction. The pile head lateral and rotational stiffness is highly nonlinear.

3.9.1 Single pile stiffness


We need the stiffness of the pile head to vertical, lateral and moment loading; in addition we need to
be able to estimate the response of the pile shaft to soil movements, which occur in kinematic
interaction.
The linear response is well documented. Two models are commonly used: 1) the elastic continuum
and 2) the Winkler spring.
There is an important difference between the vertical stiffness of a pile, particularly if the length of
the pile is modest and it bears on a firm stratum, and the lateral stiffness. The axial stiffness of a
vertical pile is larger than the lateral stiffness, it may involve soil-pile interaction over the full length
of the pile shaft, whereas the lateral stiffness mobilizes a relatively short portion of the pile shaft.

Elastic Continuum Model


The deflected shape of a pile which has horizontal lateral load applied at the ground surface extends
a distance of several pile diameters into the soil profile. Beneath this there is negligible lateral
displacement.
The length over which the lateral displacement occurs is known as the active length and it is a
function of 1) pile diameter, 2) the elastic modulus of the soil, and 3) the ratio of pile modulus to soil.

3.10 SSI effects on the seismic response of structures


• SSI responses represent a more realistic behavior of structures during seismic excitation.
• SSI effects generally provide beneficial effects. These effects reduce the seismic responses in
terms of acceleration seismic forces. However, some other responses such as displacement,
the P– effects tend to increase
• The beneficial effects of SSI increase as the embedment depth of the structure increases.
• Erroneous properties may result from characterizing dynamic properties of the structural
systems from analytical models and vibration tests without due consideration to SSI effects.
• SSI alters the free-field motions, stresses and deformations in the soil mass. For this reason,
SSI effects may become beneficial for soil sites prone to liquefaction (Seed et al., 1990)
• Large structures with significant SSI effects alter the motions of ground in their proximity.
The response of nearby smaller structures can be influenced by the vibration of the large
structure.
• For partly embedded structures, maximum seismic shear moment occur at about the ground
surface evaluation.

54
EC8 Guide Book Chapter 3 – Soil–structure interaction

• Seismic soil pressures on the embedded walls of the structures are significantly affected by
the SSI effects. The dynamic inertia SSI characteristics of the structure control the magnitude
and the distribution of the soil pressure. Seismic soil pressure between two adjacent
structures may increase significantly due to through-soil structure –to-structure interaction
effects.
• SSI effects increase the period of the structure and/or introduce a new period of vibration.

55
EC8 Guide Book Chapter 3 – Soil–structure interaction

56
EC8 Guide Book Chapter 4 – Structural analysis

4 Structural analysis
Dominik H. Lang and Emrah Erduran (NORSAR, Kjeller)

4.1 Importance classes and importance factors


In order to account for the importance of the building under consideration the seismic input ground
motion may be either scaled up or down. This is ensured by importance factor I , which is used in
order to come from the reference peak ground acceleration agR to the design ground acceleration ag:
ag = agR  I eq. (4.1)

Buildings are classified into four importance classes, each having a distinct importance factor I,
according to the following criteria:
1. Severity of impact to human lives in case of collapse.
2. Importance for public safety and civil protection issues in the aftermath
3. Connected (social and) economic consequences in case of collapse

The importance factors γI are NDP's ('Nationally Determined Parameters') except for importance class II
(i.e. γI = 1.0), which is associated with a seismic event with a reference return period TNCR. The values
given for Norwegian conditions are more extreme than the general European average (Table 4.1).

Table 4.1 Building importance classes and their allocated importance factors as proposed for general
European conditions and Norway.

Importance I I
Buildings
class (General) (Norway)
Buildings of minor importance for public safety, e.g.
I 0.8 0.7
agricultural buildings, etc.
II Ordinary buildings, not belonging in the other categories 1.0 1.0
Buildings whose seismic resistance is of importance in view of
III the consequences associated with a collapse, e.g. schools, 1.2 1.4
assembly halls, cultural institutions etc.
Buildings whose integrity during earthquakes is of vital
IV importance for civil protection, e.g. hospitals, fire stations, 1.4 2.0
power plants, etc.

57
EC8 Guide Book Chapter 4 – Structural analysis

4.2 Methods of analysis proposed by EC8


4.2.1 Linear vs. non-linear analysis

The design of a building assuming to remain elastic during a large earthquake


may be conservative, however, also uneconomic as the expected seismic
demands are overestimated!

This means that ductility of a structure should preferably be considered in order to reduce the
seismic demands to acceptable (i.e. more economical) levels.

4.2.2 Ductility  and behavior factor q


Ductility is generally defined as “the ability of a building or element to withstand large deformations
beyond its yield point without fracture (here: collapse)”. The dimensionless ratio between maximum
displacement umax and yield displacement uy is called ductility factor  :
u max
 eq. (4.2)
uy

To design for ductility means, e.g.:


→ plastic hinges form in beams before they do in columns,
→ adequate confinement to concrete using closely spaced hoops,
→ ensure that steel members fail away from connections,
→ avoid large irregularities,
→ in RC members, shear strengths are significantly higher than flexural strengths.

‘Yielding’ of a structure means limiting of the peak force that it must sustain; this reduction is
represented by behavior factor q:
F0 u 0
q  eq. (4.3)
Fy u y

F
F0 = q  Fy corresponding linear (elastic)

elastoplastic system
Fy

uy uo umax =   uy u

Figure 4.1. How ductility factor  and behavior factor q are related.

58
EC8 Guide Book Chapter 4 – Structural analysis

4.2.3 Building design vs. analysis method


Eurocode 8 proposes the use of four different analysis methods. These methods will be discussed in
more detail in the subsequent subchapters:
a) Linear-elastic methods
(1) Lateral force method of analysis
(2) Modal response spectrum analysis
b) Non-linear (inelastic) methods
(3) Non-linear static ('pushover') analysis
(4) Non-linear time history analysis.

Depending on the level or regularity of the building’s design, Eurocode 8 allows the use of simplified
analysis methods (Table 4.2).

Table 4.2 Relation between structural regularity and seismic analysis and design.

Regular in plan Regular in elevation Building model Allowed linear-elastic analysis method

● ● lateral force (4.2.4)


planar (2D)
● ○ modal (4.2.5)

○ ● lateral force (4.2.4)


spatial (3D)
○ ○ modal (4.2.5)

4.2.4 Lateral force method


The lateral force method shall be generally applied to buildings:
- whose response is dominated by the first mode; which means that their natural period of vibration
T1 is limited to a certain factor:

 4  TC
T1   eq. (4.4)
2.0 sec
- that are regular in elevation (see Table 6.2).

Base shear force Fb


a) Calculation of fundamental period of vibration T1:
- based on any equation coming from structural mechanics (e.g. Rayleigh method)
- for building heights H  40 m :

T1  C t  H 0.75 eq. (4.5)


with: H - building height (in [m]) from foundation or top of a rigid basement
Ct - structural coefficient, i.e.
Ct = 0.085 for moment-resistant steel frames

59
EC8 Guide Book Chapter 4 – Structural analysis

0.075 for moment-resistant concrete frames and eccentrically braced


steel frames
0.050 for all other structures
Ct = 0.075 / √Ac for building with concrete or masonry shear walls
with Ac acc. to eq. (7.y)


Ac   Ai  ( 0.2  ( l wi / H )) 2  eq. (4.6)
with: Ac - total effective area of the shear walls in the first storey (in [m2])
Ai - effective cross-sectional area of shear wall i (in [m2])
lwi - length of shear wall i parallel to applied forces

b) Base shear force Fb:


- shall be calculated for each horizontal direction
Fb  Sd ( T1 )  m  
eq. (4.7)
with: Sd (T1) - ordinate of the design spectrum at T1
m- total mass of the building
- correction factor (0.85 if T1  2TC and if N > 2, otherwise  = 1.0)3

Distribution of horizontal seismic forces


a) Calculation of lateral seismic forces for each story level:
- calculation of horizontal forces Fi to all storey levels can be done by two ways:
Type A (dependent on height of masses):
z i  mi
Fi  Fb 
z j  m j eq. (4.8)
with: zi  height of the respective mass i

Type B (dependent on absolute horizontal displacement of masses):


s i  mi
Fi  Fb 
s j m j eq. (4.9)
with: si  lateral displacement of mass i in the 1st mode

b) Calculation of story masses (seismic masses) mi :


- composed of permanent Gk and participating live loads Qki :
Gk + ∑i (ΨEi  QKi) eq. (4.10)
with: ΨEi - combination coefficient for variable action
Ψ2i - occupancy type coefficient (see EN 1990:2002)
φ- load type coefficient (EC8, 4.2.4)
while: ΨEi = φ  Ψ2i

3 For buildings with more than 2 stories the effective modal mass for the 1st mode is smaller, in fact reduced by 15% on average.

60
EC8 Guide Book Chapter 4 – Structural analysis

Tutorial on lateral force method:

Tutorial 4.1: Lateral force method for a 3-story RC frame building (residential use)

Given parameters: m3
behavior factor q=4 m2
ground motion agR = 0.3 g 3 x 3.5 m
RC frame  Ct = 0.075 m1
residential use  γI = 1.0
 ΨEi = φ  Ψ2i = 0.8  0.3 = 0.24
Ground type C
Ground type C  S = 1.15
(Type 1)  TB = 0.2, TC = 0.6, TD = 2.0
Seismic masses:
Level G [kN] Q [kN] G+ Ψ  Q [kN] Mass mi [tons]
3 260 120 289 29.44
2 350 140 384 39.10
1 750 300 822 83.79
Total seismic mass m 152.33

Base shear force Fb:


1. Fundamental period T1 :
T1 = Ct  H0.75 = 0.075  10.50.75 = 0.44 s
Criteria check: T1 = 0.44 s  4  TC = 4  0.6 s = 2.4 s 
T1 = 0.44 s  2.0 s 
2. Design spectral acceleration:

ag = agR  I = 2.943  1.0 = 2.943 m/s2

Sa,d = ag  S  2.5/q = 2.943  1.15  2.5/4.0 = 2.12 m/s2

3. Base shear force Fb :


since T1 < 2  TC →  = 0.85
Fb = Sa,d (T1)  m   = 2.12  152.33  0.85 = 274 kN
Load distribution and moment calculation:
Level Height z Mass mk zk  mk Force Fk Moment = Fk  zk
3 10.5 m 29.44 tons 309.12 mtons 96.68 kN 1015.1 kNm
2 7.0 m 39.10 tons 273.70 mtons 85.60 kN 599.2 kNm
1 3.5 m 83.79 tons 293.27 mtons 91.72 kN 321.0 kNm
Totals 152.33 tons 876.09 mtons 274.0 kN 1935.3 kNm

61
EC8 Guide Book Chapter 4 – Structural analysis

Tutorial 4.1: Lateral force method for a 3-story RC frame building (residential use)

Effective height of the resultant lateral force


M res 1935.3 F3
heff    7.06 m m3
Fres 274.0 Fres
m2
F2
heff
m1
F1
Mres

4.2.5 Modal response spectrum method


The modal response spectrum method shall be applied if the criteria for the lateral force method are
not fulfilled. This generally applies to taller buildings (with some stories) for which the following
criteria apply:

 4  TC
T1   eq. (4.11)
2.0 sec
The main difference to the lateral force method consists in the fact that higher modes are to be
considered. The question how many and up to which mode shall be included is specified by the
following criteria:
(1) response of all modes shall be considered that contribute significantly to the global building
response (i.e. important for buildings of a certain height),
(2) those modes shall be considered for which:
a. the sum of their modal masses is at least 90% of the total mass, i.e. mi ≥ 0.9  mtot
b. their modal mass is larger than 5% of the total building mass, i.e. mi ≥ 0.05  mtot

Criteria (2a and (2b) may not be fulfilled for buildings with lower torsional modes (i.e. buildings that
are prone to torsional effects). In this case the maximum mode number k shall be defined by both of
the following criteria:
k ≥ 3  √n eq. (4.12)
Tk ≤ 0.20 sec eq. (4.13)
with: k – maximum mode number
n – story number (from top of foundation to building top
Tk – period of vibration of mode number k

Modes may be combined if their periods are very close to each other. The condition for independent
mode shapes is fulfilled if the periods Ti and Tj of neighboring mode shapes i and j are:

62
EC8 Guide Book Chapter 4 – Structural analysis

Tj ≤ 0.9  Ti eq. (4.14)


with: Ti > Tj

The modal response spectrum method is based upon solving the differential equation:
M  u  C  u  K  u  0 eq. (4.15)
However, when using structural analysis software the required parameters are directly given, i.e.
periods Ti and mode shapes i .

  j ,1 
n,1 n,2 n,3  
1   j 1,1 
 
 n ,1 
j+1,1 j+1,2 j+1,3
  j ,2 
 
j,1 j,2 j,3 2    j 1, 2 
 
 n,2 
  j ,3 
 
T1 T2 T3 3    j 1,3 
 
 n ,3 
Figure 4.2 Mode shapes and corresponding mode shape vectors.

Lateral story loads Fj,i :


a) Calculation of modal participation factors i :
With the story masses m and relative mode shape amplitudes i the modal participation factors can
be derived:
n

m
j 1
j   j ,i
i  n

m
j 1
j   2j ,i
eq. (4.16)
b) Spectral accelerations Sa(Ti) for each mode i:
Take the spectral accelerations from the design spectrum at the respective period Ti.

63
EC8 Guide Book Chapter 4 – Structural analysis

Spectral acceleration Sa
Sa(T3)

Sa(T2)
Sa(T1)

T1 T2 T3
Period T [sec]
Figure 4.3 Deriving the design spectral accelerations Sa for each mode shape with period Ti .

c) Lateral story loads Fj,i :


F j ,i  m j   j ,i   i  Sa (Ti )
eq. (4.17)
Maximum shear forces Fb and moments M :
If the response of more independent mode shapes have to be considered, the maximum value of a
seismic force F, moment M, displacement u, etc. may be derived by the following equations.
a) Shear forces Fb :
n
Fb , m  
i 1
Fb2, m , i eq. (4.18)

b) Moments M :
n
Mm  
i 1
M m2 , i eq. (4.19)

64
EC8 Guide Book Chapter 4 – Structural analysis

Tutorial on Modal response spectrum method (following an example of Betonkalender 1978):

Tutorial 4.2: Modal response spectrum method for a 3-story RC frame building

Given parameters: m3 = m
behavior factor: q=4 h k3 = k
ground motion: agR = 0.3 g m2 = 1.5m
h k2 = 2k
residential use: γI = 1.0
m1 = 2m
E = 2.1  10 kN/m
8 2
h k1 = 3k
I = 2.679  10-5 m4
m = 50 tons = 50 kNs2/m
h = 3.0 m
Ground type C (Type 1)  S = 1.15, TB = 0.2 s, TC = 0.6 s, TD = 2.0 s
k = 12  EI/h3

Setting up the differential equation of motion:


M  u  C  u  K  u  0 Assumption: Damping matrix [C] = 0
Mass matrix: Stiffness matrix:
 m1 0 0  2 0 0  k1  k2  k2 0   5 2 0 
       
M    0 m2 0   m   0 1. 5 0  K     k2 k 2  k3  k3   k    2 3  1
0 m 3  0 0 1  0  k3 k3   0 1 1 
 0     
Modal segmentation:

K    2  M   0 5k  2m 2  2k 0
 2k 3k  1.5m 2 k 0
0 k k  m 2

Modal circular frequencies i and periods Ti :

1 = 4.19 s-1 → T1 = 1.50 sec


2 = 8.97 s-1 → T2 = 0.70 sec
3 = 13.3 s-1 → T3 = 0.47 sec
Modes i :

 0.30    0.676   2.47 


     
1   0.644  2     0.601  3    2.57 
 1.00   1.00   1.00 
     

65
EC8 Guide Book Chapter 4 – Structural analysis

Tutorial 4.2: Modal response spectrum method for a 3-story RC frame building

n,1 n,2 n,3

j+1,1 j+1,2 j+1,3

j,1 j,2 j,3

T1 T2 T3

Modal participation factors i :


n

m
j 1
j   j ,i
i
i  n

Mi *
m
j 1
j   2j , i

1 = 100  0.3 + 75  0.644 + 50  1.0 = 128.3 kNs2/m


2 = –100  0.676 – 75  0.601 + 50  1.0 = -62.7 kNs2/m
3 = 100  2.47 – 75  2.57 + 50  1.0 = 104.3 kNs2/m

M1* = 100  0.32 + 75  0.6442 + 50  1.02 = 90.0 kNs2/m


M2* = 100  0.6762 + 75  0.6012 + 50  1.02 = 122.8 kNs2/m
M3* = 100  2.472 + 75  2.572 + 50  1.02 = 1155.0 kNs2/m

1 = 128.3 / 90.0 = 1.426


2 = -62.7 / 122.8 = –0.511
3 = 104.3 / 1155.0 = 0.090
Design spectral accelerations Sa(Ti ) for each mode i :

2.5  TC  2.5  0.6 


T1 = 1.50 sec : Sa ( T )  a g  S      ( 2.943 1.0 ) 1.15     0.846 m / s 2
q  T1  4.0 1.50 
Check: Sa(T) = 0.846 m/s2 ≥ β · ag = 0.20· 2.943 = 0.5886 m/s2 

2.5  TC  2.5  0.6 


T2 = 0.70 sec : Sa ( T )  a g  S      ( 2.943 1.0 ) 1.15    1.813 m / s 2
q  T1  4.0  0.7 
Check: Sa(T) = 1.813 m/s2 ≥ β · ag = 0.20· 2.943 = 0.5886 m/s2 
2.5 2.5
T3 = 0.47 sec : S a (T )  a g  S   ( 2.943 1.0 ) 1.15   2.115 m / s 2
q 4.0

66
EC8 Guide Book Chapter 4 – Structural analysis

Tutorial 4.2: Modal response spectrum method for a 3-story RC frame building

Lateral story loads Fj,i :

F j ,i  m j   j ,i   i  S a ( Ti )

F1,1 = 100  0.30  1.426  0.846 = 36.2 kN


F2,1 = 75  0.644  1.426  0.846 = 58.3 kN
F3,1 = 50  1.00  1.426  0.846 = 60.3 kN
F1,2 = 100  (–0.676)  (–0.511)  1.813 = 62.6 kN
F2,2 = 75  (–0.601)  (–0.511)  1.813 = 41.8 kN
F3,2 = 50  1.00  (–0.511)  1.813 = –46.3 kN
F1,3 = 100  2.47  0.090  2.115 = 47.0 kN
F2,3 = 75  (–2.57)  0.090  2.115 = –36.7 kN
F3,3 = 50  1.00  0.090  2.115 = 9.5 kN

Mode 1: Mode 2: Mode 3:

F3,1= 60.3 F3,2 = –46.3 F3,3 = 9.5

F2,1 = 58.3 F2,2 = 41.8 F2,3 = –36.7

F1,1 = 36.2 F1,2 = 62.6 F1,3 = 47.0

Maximum shear forces Fb :

Mode 1: Mode 2: Mode 3: Resulting shear force:

60.3 -46.3 9.5 76.6

118.6 -4.5 -27.2  121.7

154.8 58.1 19.8 166.5

n
with: Fb , m  
i 1
Fb2, m , i

67
EC8 Guide Book Chapter 4 – Structural analysis

4.2.6 Non-linear static ('pushover') analysis


A broad variety of analysis methods both elastic (linear) and inelastic (nonlinear), are available for the
design of future buildings or for checking structures that already exist. Since the most customary
inelastic analysis procedure insists in the nonlinear time history analysis, which is regarded as
impractical and time-consuming, simplified nonlinear analysis procedures provide a more manageable
tool. Even though nonlinear time history analyses are considered to provide more exact results,
simplified nonlinear procedures are believed to produce satisfactory results for many particular tasks.
For many purposes in earthquake engineering and seismic risk assessment, the pushover analysis
establishes a main component, e.g. when applying the capacity spectrum method (CSM). Here, the
outcome of the nonlinear static pushover analysis, the capacity curve of the structure under
investigation is combined with a reduced (critically damped) response spectrum representing the
(damaging) seismic excitation or the seismic demand. By superposing the structural capacity and
seismic action (demand), the performance of the structure under lateral earthquake impact can be
assessed (ATC, 1996; Fajfar, 2000; Freeman, 1998; Freeman et al., 1975).

The ‘pushover’ analysis is a non-linear static analysis method that is carried out under conditions of
constant gravity loads and monotonically increasing horizontal loads. Pushover analysis may be
applied to verify the structural performance of newly designed buildings for the following purposes:
(1) to verify or revise the overstrength ratio values u
(2) to estimate the expected plastic mechanisms and the distribution of damage,
(3) to assess the structural performance of existing or retrofitted buildings (EN 1998-3),
(4) as an alternative to the design based on linear-elastic analysis methods, which use behavior
factor q.

Depending on the level or regularity of the building’s design, EC8 allows the use of simplified
modeling assumptions (Table 4.3).

Table 4.3 Relation between structural regularity and modeling assumptions for pushover analysis.

Regular in plan Regular in elevation Building model

● ●
planar (2D)
● ○
○ ●
spatial (3D)
○ ○

The major outcome of a pushover analysis is a ‘capacity curve’ that displays the base shear – roof
displacement relationship of a structure (Figure 4.4).

68
EC8 Guide Book Chapter 4 – Structural analysis

d
V

Base shear V
Roof displacement d

Figure 4.4. Deriving the design spectral accelerations Sa for each mode shape with period Ti .

Step-by-Step Pushover Analysis:


1. Create a non-linear computer model of the building.
Non-linear behaviour of bending members is modelled by the corresponding moment-
curvature relationship over a finite length; plastic hinge length lpl. The plastic hinge length
designates the length of the zone where the non-linear action extends towards the middle of
the element from the face joint. For reinforced concrete members lpl is generally assumed to
be between 0.5d and 1.0d (with d – depth of the member).

Moment

lpl lpl
i j

plastic hinge region Curvature

Figure 4.5. Definition of plastic hinge length lpl and moment curvature relations.

69
EC8 Guide Book Chapter 4 – Structural analysis

2. Apply gravity analysis considering dead loads and reduced live loads.

G + ΨEi  Q

Base shear V
Roof displacement d

Figure 4.6. Application of gravity loads.

3. Apply lateral loads to the structure in proportion to the selected load pattern.
d
V
Base shear V

Roof displacement d

Figure 4.7. Application of lateral load pattern.

70
EC8 Guide Book Chapter 4 – Structural analysis

Table 4.4 Types of load pattern for pushover analysis.

In general, load patterns that are used in pushover analysis can be categorized into five types:

Type 1: Uniform Load Pattern: Lateral loads at the


story level are proportional to the story mass.
Fx = [mx /∑mx ]·V eq. (4.20)

Type 2: Triangular Load Pattern: Lateral loads at the


m3
story level are proportional to the product of story mass F3
and story height. m2
Fx = [mx hx/∑mx hx]·V eq. (4.21) F2
m1
Type 3: First Mode Load Pattern: Lateral loads at the F1
story level are proportional to the product of story mass
and the first mode shape.
Fx = [mx x/∑mx x]·V eq. (4.22)

Type 4: Adaptive Load Pattern: same as Type 3 until


first yielding; the mode shapes are adapted after first
yielding.

Type 5: Multi-modal Load Pattern: V

The effects of higher modes are taken into account in


the load pattern.

nth mode 1st mode

According to EN 1998-1, at a minimum, load patterns 1 and 3 needs to be considered in non-linear


static analysis.

71
EC8 Guide Book Chapter 4 – Structural analysis

4. Calculate member forces under the applied lateral load and gravity load combination and
check for yielding in the members. Record the base shear V and roof displacement d.
d
V

Base shear V
Roof displacement d

d
V

Base shear V

Roof displacement d

Figure 4.8. Calculation of member forces for different analysis steps until a mechanism forms.

72
EC8 Guide Book Chapter 4 – Structural analysis

Tutorial on pushover analysis:

Tutorial 4.3: Pushover analysis for a portal frame with a rigid beam hinged to the columns

Structure: Portal frame with a rigid beam hinged to the columns

F My = 450 kNm
M
Ky = 0.009 rad/m
2
1 2 5.0 m
1
My = 250 kNm
Ky = 0.0125 rad/m

5.0 m 
: hinge

Deformed shape: Free body diagram:


F   V1 V2

M1 M2
V1 V2

Equilibrium equations:
V1 + V2 = F
M1 = 5 · V1
M2 = 5 · V2

Elastic range:
In the elastic range, the deformation demand of the frame can be computed by taking the moment of
the curvature diagram about the top of the columns (using moment-area theorem).

Vi EI1 = My1/y1 = 20,000 kNm2


EI2 = My2/y2 = 50,000 kNm2
2/3 h
h from column :
x
δ = M1 · 5 · 2/3 · 5/20,000 = 8.33 e-4 · M1
Mi
from column :
Vi Mi i = Mi/EIi δ = M2 · 5 · 2/3 · 5/50,000 = 3.33 e-4 · M2

73
EC8 Guide Book Chapter 4 – Structural analysis

Tutorial 4.3: Pushover analysis for a portal frame with a rigid beam hinged to the columns

Since the displacements at the top of each column are equal: M2/M1 = 2.5

Using this relationship, at the instant of yielding of the second column, i.e. M2 =450 kNm, the
moment demand in the first column will be M1 = 180 kNm < My2 . Therefore, first yielding will occur
at the base of column .

At the instant of first yielding in the system:


V1 = M1/5 = 36 kN
V2 = M2/5 = 90 kN
F = V1 + V2 = 126 kN
δ = 8.33 e-4 · M1 = 0.15 m

Inelastic Range:
At this point, plastic hinges form at the bottom of column . The moment demand in the plastic
hinge region cannot increase further:
M2 = 450 kNm => V2 = 90 kN

However, the first column is still in the elastic range. In the next step, the first column will also reach
the yielding point:
M1 = 250 kNm => V1 = 50 kN

F = V1 + V2 = 140 kN

The displacement demand can be computed using the relationship for column :
δ = 8.33 e-4 · M1 = 0.208 m

Plastic Mechanism:
Once plastic hinges form at the bottom of both columns, the structure becomes a plastic mechanism.
Moment demands, and in turn shear force demands, can increase in neither column. Therefore, the
base shear force remains constant while the displacement demands increase. The resulting capacity
curve is given as:
160
140
120
F = 140 kN
F = 126 kN
Base Shear, F (kN)

= 0.208 m
100 = 0.150 m
80
60
40
20
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Roof Displacement ,  (m)

74
EC8 Guide Book Chapter 4 – Structural analysis

Determination of Target Displacement:


1. The capacity curve obtained from pushover analysis of the MDOF system is converted to the
capacity curve of a SDOF system
a. Compute the first mode participation factor 
n

m
j 1
j   j ,1
 n

m
j 1
j   j2,1 eq. (4.23)

b. Compute the base shear V* and displacement d* of the equivalent SDOF from V and
d (base shear and displacement of the MDOF system):

V
V* 
 eq. (4.24)
d
d* 
 eq. (4.25)
c. The mass of the equivalent SDOF is computed as:
m *   m i   j ,1 eq. (4.26)
d. The capacity curve of the SDOF system is then bi-linearized:
V*
A  E* 
Vy* d y *  2   d m*  m*  eq. (4.27)
 F y 

Em*– actual deformation energy up to the
formation of the plastic mechanism

dy* dm* d*

e. The period of the idealized equivalent SDOF system T* can then be determined as:

m * d *y
T  2
*

Fy*
eq. (4.28)
f. The target displacement of the elastic system det* with period T* is computed as:
2
T * 
d et *
 Se ( T )   
*

 2  eq. (4.29)

75
EC8 Guide Book Chapter 4 – Structural analysis

g. The target displacement of the inelastic system dt* is computed using different
expressions for structures in short-period range and for structures in the medium-
and long-period ranges:
i. if T* < TC (short-period range):
if Fy*/m* ≥ Se (T* ), the response is elastic and thus
dt* = det*
if Fy*/m* < Se (T* )
d et*  T 
d t   1  q u  1  c*   d et*
*

qu  T 
where: qu – ratio between the acceleration in the structure with
unlimited elastic behavior Se(T*) and in the structure with
limited strength Fy*/m* :
Se ( T * )  m *
qu 
F y*
ii. if T* ≥ TC (medium- and long-period range):
dt* = det*

76
EC8 Guide Book Chapter 4 – Structural analysis

Tutorial on target displacement demand of an inelastic system:

Tutorial 4.4: Computation of target displacement demand of an inelastic system

Design spectrum parameters:

Spectral acceleration Sa (m/s2)


ag = 0.32 m/s2 =5%
S = 1.40 1.12
TB = 0.15 s
TC = 0.35 s
TD = 1.50 s
0.45

0.15 0.35 1.50


Period T
V*
Structural parameters:

Vy* = 1180.0 kN Vy*


m* = 1250.0 t
dy* = 0.002 m

dy* d*

Effective period and elastic spectral acceleration:

m *  d *y (1250.0 )  ( 0.002 )
T  2
*
*
 T *  2  T *  0.29 s  TC
F y (1180.0 )
Se(T* ) = 1.12 m/s2

Target displacement of the elastic system:


2
T * 
2
 0.29 
d et *
 e
S ( T )  
*
 d et *  (1.12 )    d et *  0.0082 m
 2   2 

Target displacement of the non-linear system:

Fy*/m* = (1118.0/1250.0) = 0.94 < Se (T*)


Se (T * )  m * (1.12 )  (1250.0 )
qu   qu   q u  1.19
F y* (1180.0 )
d et*  T  0.0082  0.35 
dt *  1  q u  1 c*   d t  1  1.19  1   d t  0.0085 m
* *

qu  T  1.19  0.29 

77
EC8 Guide Book Chapter 4 – Structural analysis

4.2.7 Non-linear time history analysis


Time history analyses are based upon accelerograms that are applied at the building’s base. The
calculation of the structure’s performance under seismic action can be conducted either assuming
elastic or inelastic behavior. For the non-linear time history analysis, as suggested by EN1998:1.2004,
the mathematical model shall include the strength of structural elements as well as their post-elastic
behavior. In case that realistic models are available, the non-linear time history analysis is definitely
the most accurate methodology: structural behavior including damage progression effects can be
realistically traced, which allows an optimized structural design.
Time history analyses are based upon time-dependent numerical procedures, which are generally very
complex and require powerful calculating capacities. In order to allow statistically secured results, a
larger number of accelerograms should be used. At the same time, these accelerograms should be
representative for the respective building site, e.g. in terms of soil conditions, distance to the source
etc. Since this prerequisite is often not fulfilled, it is allowed to use simulated (synthetically generated)
accelerograms that are consistent with source and path mechanisms and the underlying soil
conditions. Accelerograms shall be chosen according to the provisions given in EN 1998-1, Section
3.2.3.1.

Step-by-step non-linear time history analysis:


1. Create a non-linear computer model of the building. Non-linear cyclic behaviour of bending
members is modelled by the corresponding moment-curvature relationship over a finite
length; plastic hinge length lpl .
Moment

lpl lpl
i j

plastic hinge region Curvature

Figure 4.9. Definition of plastic hinge length lpl and moment curvature relations.

78
EC8 Guide Book Chapter 4 – Structural analysis

2. Apply gravity analysis considering dead loads and reduced live loads.

G + ΨEi  Q

Figure 4.10. Application of gravity loads.

3. Solve equation of motion at each time step and record the response parameters:
.
MÜ  C U  KU   MIü g eq. (12)

Determination of Response Parameters:


If the response is obtained from at least 7 non-linear response history analysis under different
accelerograms, the average of the response quantities from all of these analyses should be used as the
design value of the action effect Ed . Otherwise, the most unfavorable value of the response quantity
among the analyses should be used as Ed .

79
EC8 Guide Book Chapter 4 – Structural analysis

80
EC8 Guide Book Chapter 5 – Design of concrete buildings

5 Design of concrete buildings


Anton Gjørven (Norconsult, Sandvika)

5.1 Tutorial 5.1: Building based on precast concrete elements

Figure 5.1 Precast element building.

Assume:
- Seismic class III
- Ground Type C
- ag,40Hz = 0.55 m/s2
- Building layout: height h = 19.0 m
Precast elements
Gyration radius > stiffness radius
Upper 50% of the mass has its centre at el. 13
The building is regular in elevation.
The building is above ground.
Wind exposed area 400 m2
Form factor wind = 2.0

Proposed solution:
- Soil parameters: S =1.4 , TB = 0.15 s, TC = 0.35 s, TD = 1.50 s (middle dense sand)
- Building importance: γI =1.4 (institutional occupancy; see NA.4.2.5, Table NA.4(901))
- see NA.3.2.1(4):

81
EC8 Guide Book Chapter 5 – Design of concrete buildings

ag · S = 1.4 · 0.8 · 0.55 m/s2 · 1.4 = 0.862 m/s2 < 0.98 m/s2 → DCL can be used if q ≤ 1.5 !
Note: Since ag · S < 0.98 m/s2 it is sufficient to check base shear compared with wind and soil
pressure (soil pressure = 0 in this case), see also 4.4.1(2)!

- Classification of the building:


1) Due to gyration radius greater than stiffness radius, the system must either be calculated as
a torsionally flexible system or an inverted pendulum system (see 5.2.2.1(4)P and (6)). This is
to ensure a peripheral and well-distributed stiffness performance in ductile design. The
present type of building, with a central concrete core, may not have this property.

2) Choices: to be able to use DCL: q ≤ 1.5. This means that the building may be designed as
an inverted pendulum. However, if the centre of gravity of 50% of the upper part mass is
located at the upper third of the height, the building must be considered as a torsionally
flexible system, which gives the demand to q : q ≤ 2 (and DCM), which in turn may gives a
smaller base shear. Note that material factors for concrete and steel increase when going
from DCL to DCM (see NA. 5.2.4(3)). For precast elements it could be of significance that
the material factor for steel increases from 1.0 to 1.15 (by a factor of 15%). Note further that
the minimum amount of reinforcement in precast elements shall be 1%.

3) Ductility of precast members - energy absorption:


A kP factor of 0.5 has to be introduced, unless it can be shown that connections behave
ductile. q = kP · q0 = 0.75 (see NA 5.11.1.3.2).
In this example, we choose to base the precast element design on an overdesigned
connection and connections located away from critical regions (see 5.11.2.1.1 and 5.11.2.1.2,
and Figure 5.2a and b). We do then have to use an overstrength factor of 1.2.

The Figure 5.2 shows also a proposal for solution to the connection. On the right side of the
figure is shown an example of element used in buildings from one Norwegian vendor. It will
be necessary to show also the way they will connect elements. A proposal is shown on the
left side. The load transfer from the rebar to the concrete element is held outside the
connection point, compare with a) and b).

82
EC8 Guide Book Chapter 5 – Design of concrete buildings

f)
e)
Figure 5.2.

- Calculation of base shear:


h = 19.0 m: T1 = Ct  H0.75 = 0.075  19.00.75 = 0.683 s
f1 = 1/T1 = 1.465 Hz

f3 (FEM)= 1.5354 Hz
T3 (FEM)= 1/f3(FEM) = 0.651 s

Sd (T1) = (0.8 · ag,40Hz · γI) · S · 2.5/q · (TC/T) = 0.884 m/s2


Sd (T1) = (0.8 · 0.55 · 1.4) · 1.4 · 2.5/1.25 · (0.35 / 0.683) = 0.884 m/s2
Sd (T3 (FEM)) = (0.8 · 0.55 · 1.4) · 1.4 · 2.5/1.25 · (0.35 / 0.651) = 0.927 m/s2

Fb = m · Sd (T1) = 1,256,555 kg · 0.884 m/s2 = 1,111 kN


from analysis: Fb = 1,099 kN

Since ag · S is less than 0.1g we need no further dynamic seismic calculation of the building
than the base shear calculation. In this tutorial case, there is, however, performed an FEM-
analysis to check the stresses in the connections. The normal case would be to put on an
equivalent wind load in a static case and calculate the forces in the connections based on this
static analysis. If ag · S is greater than 0.1g, however, a complete dynamic linear or non-linear
analysis shall be performed (see NA. 3.2.1(4)). This is due to the fact that the seismic impact
may be higher at higher levels in the building. This is also the case in the actual example, see
contour plots on next pages.
The base shear has to be compared with wind forces, which is the only other load case with
horizontal components. 111 tons seismic load divided on 400 m2 gives an equivalent wind load
of 0.14 tons/m2 with a form factor of 2 (see 4.4.1.(2 a))

83
EC8 Guide Book Chapter 5 – Design of concrete buildings

- Results from the analysis:


Figures 5.3 and 5.4, which are from the Response Spectrum FEM-analysis, show that a
membrane connection capacity in the order of 100 kN/m can withstand the seismic load
almost everywhere in this building. There are, however, some areas with gray spots which
means that the connection shown in Figure 5.2 is not sufficient (100 kN/m taken by 2Ø25
each element, 2.5 m width).

Comment: Since we are in an area with ag · S < 0.98 m/s2 it is not required to run an FEM-
analysis if the building is designed for other loads (i.e. wind), at a magnitude which give
greater base shear than seismic load. However, the example shows that it is extremely useful
to run such analyses to check frequencies of the building, the load at the connections etc.

Figure 5.3. Membrane stress resultants SF1.

Figure 5.4. Membrane stress resultants SF2.

84
EC8 Guide Book Chapter 5 – Design of concrete buildings

Figure 5.5. Moment SM1.

Figure 5.6. Max principal stresses.

85
EC8 Guide Book Chapter 5 – Design of concrete buildings

86
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

6 Geotechnical aspects and foundations


Farzin Shahrokhi (Multiconsult, Oslo)

6.1 Introduction
Geotechnical aspects of seismic design are described in Eurocode 8 Part 5 (EN1998-5:2004), which
deals with dynamic soil parameters, siting criteria, SSI, foundations, retaining structures, liquefaction
hazard etc.
It is strongly suggested to involve a geotechnical earthquake expert in an early phase of design to
address the critical elements of geotechnical design, and discuss the choosing of foundation concept.
The seismic site classification (soil amplification factor), evaluation of liquefaction potential, soil
degradation, defining the foundation impedance functions (spring stiffness and dashpot coefficients)
and dynamic capacity of foundations shall be evaluated by a geotechnical earthquake engineer.

6.2 Geotechnical earthquake engineering


Geotechnical earthquake engineering is defined as that subspecialty within the field of geotechnical
engineering, which deals with the design and construction of projects in order to resist the effects of
earthquakes. For geotechnical earthquake resistance design, the following tasks may need to be
performed:
 Determining the seismic action or design earthquake (EC8-1, chapter 3; EC8-5, chapter 2).
 Determining the type of foundation concept which is best suited for resisting the effects of
the design earthquake (EC8-5, chapters 4 and 5).
 Evaluating the potential of liquefaction hazard at the construction site (EC8-5, 4.1.4).
 Calculating the seismic induced settlements on the structure (EC8-5, 4.1.5).
 Defining the design parameters for the foundation, such as the bearing capacity and allowable
soil bearing pressures (EC8-5, 5.4).
 Investigating the stability of slopes for the additional forces imposed during the design
earthquake (EC8-5, 4.1.3).
 Evaluating the effect of the design earthquake on the stability of retaining structures (EC8-5,
Chapter 7).
 Calculating the foundation impedance function as an input to seismic soil – foundation -
structure interaction analysis (SFSI) of the structure (EC8-5, chapter 6).
 Developing site improvement techniques to mitigate the effects of the earthquake.

87
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

6.3 Important common definitions


‘Ground’ versus ‘soil’
In EN1998-5:2004, the term ‘ground’ is used according to the definition in EN1997-1 clause 1.5.2.3,
as “soil, rock and fill in place prior to the execution of the construction works”. Thus, a classification of ground
types is introduced in EN1998-1:2004 for the purpose of establishing the dependence of the seismic
action on the geotechnical characteristics of the construction site.

‘Shallow’ versus ‘spread’ foundations


In this case, separate definitions are used in the two codes: in EN1998-5:2004, “shallow foundations”
include footings and raft foundations, while in EN1997-1 the same are called “spread” foundations,
and include pads (isolated footings), strips and rafts.

6.4 Ultimate limit state (ULS) and damage limitation state (DLS)
Safety verifications in EN1998-5:2004 address ULSs, i.e. limit states of rupture or excessive
deformation in the ground (GEO in EN1997-1) or in structural elements (STRU in EN1997-1).
They also address damage, or serviceability limit state.
Preventing the occurrence of GEO or STRU limit states is consistent with the no-collapse
requirement set forth in EN1998-1:2004. DLSs are defined in the latter (EN1998-1:2004, clause
2.2.1) as those ‘associated with damage beyond which specific service requirements are no longer
met’.
Verification with respect to DLSs is advocated in EN1998-5 in the general requirements for:
 Slope stability (clause 4.1.3.1): it may not suffice to verify equation (Ed  Rd) to check whether
the limit states with ‘unacceptably large displacements of the ground mass’ is attained, and
actual permanent displacements of the ground mass may have to be computed during an
earthquake. Should such computations show that predicted permanent displacements are
limited and have no adverse functional effects on the structure, the slope should be
considered as safe, and its safety actually controlled by a damage limitation requirement.
 The foundation system (clause 5.1), which prescribe that seismically induced ground
deformations be compatible with the essential functional requirements of the structure.
 Earth-retaining structures (clause 7.1), for which ‘permanent displacements (..) may be
acceptable if it is shown that they are compatible with functional requirements’. Thus,
considerations similar to those introduced for slope stability apply.

6.5 Dynamic soil properties


For seismic design purposes, a series of dynamic soil parameters and properties may need to be
evaluated (EC1998-5:2004, chapter 2).
These include:
 Strength parameters:
o for cohesive soils: Undrained shear strength cu ,
o for cohesionless soils: cyclic undrained shear strength cy,u ,

88
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

o for rock: unconfined compressive strength qu ,


 Stiffness and damping parameters:
o dynamic shear modulus G,
o shear wave velocity vs ,
o damping ratio ,
o cyclic stress-strain behavior.

6.5.1 Strength parameters


Cohesive soils
Static shear strength
Earthquake loading is of short duration, and drainage cannot occur in cohesive soils during its
occurrence. The most dangerous condition is immediately after the loading, because the shear
strength will generally increase afterwards with progress f consolidation. If the development of excess
pore pressure in the soil can in first approximation be neglected, the governing shear strength will be
the same as in the static case, i.e. cu, prior to cyclic loading.

Cyclic degradation effects


When a normally consolidated specimen of cohesive soil is subjected to undrained cyclic loading,
positive u is progressively generated, and a reduction in effective stress occurs that will decrease the
shear strength and may lead to cyclic failure beyond a certain shear stress level, or certain number of
loading cycles.
The shear strength reduction induced by cyclic loading in normally consolidated (NC) clay can be
estimated by the expression
l 1
 
(c u ) cyc  1 
  eq. (6.1)
(c u ) NC  1  u  ' 
 c 

with: (cu)cyc – shear strength induced by undrained cyclic loading


(cu)NC – undrained (static) strength of the soil prior to cyclic loading
’c – effective (static) confining stress
l– experimental constant that can be estimated via the correlation l = 0.939–0.002lp
IP – plasticity index of the soil

Mentioned equation allows quantification of the undrained shear strength reduction, provided one
can estimate the cyclically induced u, which depends on the number of loading cycles.
For a typical Norwegian NC clay it is suggested to reduce the undrained shear strength as presented
in Table 1.
Table 6.1. Reduction factors for cyclic degrading of shear strength.

Importance class Reduction factor (cu)cyc / (cu)NC


I & II 85 %
III 80 %

89
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

IV 75 %

The cyclically induced decrease in shear strength is a temporary phenomenon, as drainage occurs
followed by dissipation of excess pressure, the effective stress will increase again, and the shear
strength with it, until it eventually regains, its ‘static’ value.

Rate of loading effects


In general, the rate of loading causes an increase in the strength of a cohesive soil. Considering
loading tests with times to failure of the specimen ranging from 100 s (slow test) to about 0.1 s, an
increase of the order of 15% up to 40% compared with static tests may be considered, although the
estimate is affected by the large scatter of experimental data. Such an increase is normally neglected
in favor of safety.

Cohesionless soils
In an uncemented, water-saturated sandy soil, the shear strength in undrained conditions can be
expressed by the Coulomb failure criterion in terms of effective stress:

The use of this relationship is suggested in clause 3.1(2) of EC8-5. Its use, however, requires that the
excess pore pressure, u, generated by earthquake-like cyclic loading is evaluated. This evaluation
may pose difficulties.
The alternative approach allowed in clause 3.1(1) of EC8-5 is to make use of an experimental
relationship between the undrained resistance f the soil subjected to a well-defined cyclic loading
process, denoted as cy,u, and a parameter reprehensive of the state of packing of the soil material,
such as its relative density, Dr.
Based on compilation of a set of data on typical clean sands, the relationship:

has been proposed (111), where v denotes the effective vertical stress.

6.5.2 Dynamic shear modulus Gmax


The main role played by the shear modulus G or equivalently, by the velocity of propagation of the
seismic waves vs in the ground, is in the classification of the construction site according to the ground
types established in clause 3.1.2 of EN1998-1:2004.
Additional applications that require knowledge of the shear stiffness of the soil profile include the
calculation of:
 The dynamic soil-structure interaction parameters (section 6 and Annex D of EN1998-5)
 The seismic site response, which may, for example, be necessary to define the seismic action
for ground type S1, or to obtain a refined estimate of the seismic coefficient kh to be used in
pseudo-static stability verification of earth-retaining structure higher than 10 m (clause 2 of
Annex E of EN1998-5:2004).

90
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

The dynamic shear stress Gmax is for very small shear strain levels, typically 10-5 or less. However, in
the applications just listed above, it is essential that the G values are compatible with the shear strain
levels induced by the earthquake in the ground, typically ranging between 10-5 and 10-2 if soil failure
does not occur, see section 2.2.2. Some guidance on the strain dependence of G is directly provided
in EN1998-5 (table 4.1) which is repeated Table 6.2. Alternative it can achieve by performing the
appropriate site response analysis.

Table 6.2. Average soil damping ratio and average reduction factors with in 20 m depth (EC1998-5, table 4.1).

Ground acceleration Damping ratio


ratio, .S (%)
0.10 3 0.90 (0.07) 0.80 (0.10)
0.20 6 0.70 (0.15) 0.50 (0.20)
0.30 10 0.60 (0.15) 0.36 (0.20)

Table 6.3 present some correlation for estimating dynamic shear modulus for both cohesive and
cohesionless soils.

Table 6.3. Some correlations for calculating the dynamic shear modulus.

Reference Correlation Unit Limitation

For cohesionless soils


Seed et al. (1984) kPa (K2)max = 30 for very loose sand, and
75 for very dense sands; 80 – 180 for
dense well graded gravels
For cohesive soils
Hardin (1978) kPa Pa = atmospheric pressure = 100

Mayne & Rix (1993) kPa Pa = atmospheric pressure

kPa For normal consolidated clay


with IP between 10% and 30%
For normal consolidated quick clay / sensitive clay with clay content about 40%
NGI kPa

Defining Gmax based on CPT:


CPTs are suitable for providing a detailed and continuous quantitative description of a soil profile.
CPTs are especially effective in detecting small-scale vertical (and lateral) heterogeneities.
Correlations are available between the measured cone tip resistance qc and different parameters of
geotechnical earthquake engineering such as relative density, and shear wave propagation velocity vS
or the elastic small deformation shear modulus G0. A useful correlation is:

91
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

Where qc is the measured cone tip penetration resistance (in the same units as G0), Pa is the reference
pressure (= 100 kPa), and c1 and c2 are empirical constants.

Dependency of deformation characteristics upon shear strains


The deformation characteristics of soils vary depending on the magnitude of shear strains which soils
are subjected. Overall changes in soil behavior with changes in shear strain are illustrated in Table
6.4, in which approximate ranges of the shear strain producing elastic, elasto-plastic, and failure states
of stress are indicated. In the very small shear strain level (10-5) the deformations in most soils are
purely elastic and recoverable.
Over the intermediate range of strain between 10-4 and 10-2, the behavior of soils is elasto-plastic and
produces irrecoverable permanent deformation. The development of cracks or differential settle-
ments in soil structures appears to be associated with the elasto-plastic attribute of soils within such a
range of strain.
When large shear strains exceeding a few percent, the strains tend to become considerably large
without a further increase in shear stress and failure takes place in the soils. Slides in slope or
compaction and liquefaction of cohesionless soils are associated with failure inducing large strains.

Table 6.4. Variation of soil properties with strain.

6.5.3 Shear wave velocity


The shear wave velocity of a soil is used to establish the stiffness of the soil at small strains. The
small strain (initial) shear modulus of a soil Gmax, is related to the shear wave velocity vs and the mass
density of the soil by the equation:

92
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

6.5.4 Small strain Young’s modulus

Small strain (initial) Young’s modulus Emax is related to small strain shear modulus as a function of
Poisson’s ratio  by the theory of elasticity:

For practical purposes, Poisson’s ratio of the soil can be assumed equal to 0.35 for sand and 0.45 for
clays.

6.5.5 Damping
In addition to dynamic soil–structure interaction, internal soil damping (also strain dependent
quantity) comes into play in the very same applications listed in relation to the shear stiffness.

6.5.6 Cyclic stress-strain behavior


The earthquake-induced stresses and strains that produce the most damage in soils are generally
considered to be due to cyclic shearing of the soil. The stress-strain response of soil to this type of
cyclic loading is commonly characterized by hysteresis loop. A typical hysteresis loop is shown in
Figure 6.1. The most common model used to represent the hysteresis behavior of soil in seismic
analysis is the “equivalent linear model” (Seed and Idriss, 1970). Various nonlinear constitutive
models have been also developed to represent hysteretic soil behavior (Kondner and Zelanko, 1963;
Martin, 1975; Matasovic and Vucetic, 1993).
The equivalent linear model represents nonlinear hysteretic soil behavior using an equivalent shear
modulus G equal to the slope of the line connecting the tips of the hysteresis loop and an equivalent
viscous damping ratio proportional to the enclosed area of the loop. The equivalent shear modulus
and damping ratio are strain-dependent. The strain dependence of the equivalent shear modulus and
damping ratio are described by the modulus reduction and damping ratio curves shown on Figure
6.3.

93
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

Figure 6.1. Hysteretic stress-strain response of soil Figure 6.2. Stiffness reduction with increasing strain.
subjected to cyclic loading.

Figure 6.3. Shear modulus ratio and damping ratio for hyperbolic model.

6.5.7 Depth to bedrock


Ideally, the soil profile developed for a seismic analysis should extent to competent bedrock, where
competent bedrock is defined as material with a shear wave velocity vs of at least 800 m/s, and the
physical properties of the soil over interval between the ground surface and competent bedrock
should be defined. However, if competent bedrock is not reachable at a reasonable depth, the depth
over which the physical properties of the soil for seismic design are defined should be at least 30 m.

6.6 Material factors


The recommended values of the partial factors are those given in clause NA.3.1 of
EN1998-5:2004. The partial factors γM for material properties cu , cy,u and qu are summarized as
, , and .

94
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

6.7 Slope stability


Structures adjacent to slopes (both natural and artificial) may be subjected to two different
phenomena, firstly slope instability and secondly topographic amplification.
‘Unacceptably large displacements’, mentioned in clause 4.1.3.1(2) of EN1998-5:2004, may result
from the attainment of either ULS or DLS. In either case, the seismically induced displacements in
the slope may have to be evaluated.
Structures within important class I, if the ground in static analysis is considered to be stable, the
verification of slope stability for seismic loading may be omitted (clause 4.1.3.1(3) of EN1998-
5:2004).
The design seismic action for slope stability verifications shall conform to section 2 of EN1998-
5:2004 (response spectrum compatible earthquake motions defined in clause 3.2.3.1 of EN1998-
1:2004), and the effect of topographic amplification shall be included for structure in important class
III and IV.
There are several analysis methods available including dynamic analysis using finite element or rigid
block models (Newmark method) as well as simplified pseudo-static methods.
If a refined dynamic analysis of a slope is performed by, for example, finite elements, a correct
description of strain-softening effects in the ground materials may be critical because they control the
formation of shear bands, and their development into a localized slip surface. The model should as
well account for possible effects of pore pressure built up under the cyclic loading.

6.7.1 Pseudo-static methods of analysis (Section 4.1.3 of EN1998-5:2004)


There are two limitations applied to pseudo-static method which it is essential not to overlook these
simplifications:
1. The geometry of the topographic profile and the ground profile must be reasonably regular.
2. The ground materials of the slope should not be prone to developing a significant pore
pressure increase, which may lead to loss of shear strength (loose sand) and to stiffness
degradation under cyclic loading (soft or sensitive clay), which correspond generally to site
classes S1 and S2.

The application of the pseudo-static method of analysis entails the following steps:
1. The shear strength and stiffness of the soil should be degraded as function of shear strain and
pore water pressure build up before performing the verifications as mentioned in chapter 2 of
this document and clause 4.1.3.3(2) of EN1998-5:2004.
2. After selecting a slip surface, the prescribed horizontal and vertical seismic inertia forces are
applied statistically, in addition to the other permanent loads, to the centre of the gravity of
the ground mass enclosed between the slip surface and the ground surface (clause 4.1.3.3(5)
of EN1998-5:2004).
3. Through rigid body equilibrium considerations, a safety factor is computed as the ratio of
stabilizing to destabilizing forces acting on the ground mass.
4. The same operation is repeated many times, changing the slip surface and re-computing the
safety factor, until a minimum value is found that is assumed as the effective safety factor.

95
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

The key parameter in the pseudo-static method is the seismic coefficient kh, i.e. the fraction of the
design acceleration ratio S· used to determine the design inertia forces. This fraction, set equal to
0.5 in equation (4.1) of EN1998-5:2004.

6.8 Liquefaction
6.8.1 Effect of soil liquefaction on structures
Liquefaction is a process by which non-cohesive or granular sediments below the water table lose
strength and behave as a viscous liquid when subjected to strong ground shaking during an
earthquake. Typically, saturated, poorly graded, loose, granular deposits with low fines content are
most susceptible to liquefaction.
The adverse effects of liquefaction can be summarized as follows:
 Flow failures – completely liquefied soil or blocks of intact material ride on a layer of
liquefied soil. Flows can be developed on moderate to steep slopes.
 Lateral spreads – involve lateral displacement of superficial blocks of soil as a result of
liquefaction of a subsurface layer. Spreads generally develop on gentle slopes and move
toward a free face.
 Ground oscillation – where the ground is flat or the slope too gentle to allow lateral
displacement, liquefaction at depth may disconnect overlying soils from the underlying
ground, allowing the upper soil to oscillate back and forth in the form of ground waves.
These oscillations are usually accompanied by ground fissures and fracture of rigid structures
such as pavements and pipelines.
 Loss or reduction in bearing capacity – liquefaction is induced when earthquake shaking
increases pore water pressures, which in turn causes the soil to lose its strength and hence
bearing capacity.
 Settlement – soil settlement may occur as the pore-water pressures dissipate and the soil
densifies after liquefaction. Settlement of structures may occur due to the reduction in
bearing capacity or due to the ground displacements noted above.
 Increased lateral pressure on retaining walls – occurs when the soil behind a wall liquefies and
so behaves as a “heavy” fluid with no internal friction.
 Flotation of buried structures – occurs when buried structures such as tanks and pipes
become buoyant in the liquefied soil.

6.8.2 Liquefaction potential


Section 4.1.4 of EN1998-5 describes the requirements for assessing liquefaction potential.
Furthermore it provides a normative methodology in Annex B. It should, however, be noted that
there have been numerous developments in liquefaction assessment methodologies in recent years
(e.g. Seed et al., 2003; Boulanger and Idriss, 2004).
The simplified methodology proposed by EN1998-5:2004 is based on SPT test. However, with
respect to the fact that SPT is not a usual field test in Norway, here is address other methodology
based on CPT results.

96
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

EN1998-5:2004 recommends that the shear stress approach is applied. In this method, the horizontal
shear stresses generated by the earthquake are compared with the resistance available to prevent
liquefaction. The shear stresses “demand” are expressed in terms of a cyclic stress ratio (CSR), and
the “capacity” in terms of a cyclic resistance ratio (CRR).
The CRR is assessed based on corrected CPT values using the empirically derived liquefaction charts,
which are shown in Figure 6.5.
 This expression may not be applied for depths larger than 20 m. A soil shall be considered
susceptible to liquefaction whenever CRR · CSR > l, where l is recommended to be 0.8,
which corresponds to a factor of safety of 1.25.
 If soils are found to be susceptible to liquefaction, mitigation measures such as ground
improvement and piling (to transfer loads to layers not susceptible to liquefaction), should be
considered to ensure foundation stability.
 The use of pile foundations alone should be considered with caution due to the large forces
induced in the piles by the loss of soil support in the liquefiable layers, and to the inevitable
uncertainties in determining the location and thickness of such layers.
 For buildings on shallow foundations, liquefaction evaluation may be omitted when the
saturated sandy soils are found at depths greater than 15 m.

6.8.3 Liquefaction analysis steps


Step 1 – Estimate the cyclic shear stress with depth using following equation:

where is the ratio of the design ground acceleration on type A ground to the acceleration

of gravity, g; S is the soil factor and V0 is the total overburden pressure. is depth reduction factor,
see Figure 6.4.

Figure 6.4. rd versus depth curves developed by Figure 6.5. Liquefaction assessment using corrected
Seed and Idriss (1971). CPT values.

97
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

Step 2 – Determine the cyclic shear stress ratio CSR according to:

where - vertical effective stress at the depth of interest.

Step 3 – Determine the normalized CPT value, , with depth from equation:

Step 4 – read the soil resistance to liquefaction for earthquake with M=7.5, , from
Figure 6.5. If the earthquake magnitude is less than 7.5, then the CRR should be increased according
to

The Magnitude Scaling Factor (MSF) is equal to 2.86 for Norway, ref. Table B1 of
EN1998-5:2004.

Step 5 – calculate the ratio of with depth.


The factor of safety against liquefaction is determined from the demand and resistance. A factor of at
least 1.25 is desired to prevent damages.

6.9 Foundation system


Once the inertia forces from dynamic response of structure which will be transmitted to the soil by
the foundation are determined, it must be checked that these forces can be safely supported by:
1. the foundation must not experience a bearing capacity failure
2. no excessive permanent displacements will happen
At this point a major difference appears between static, permanently acting loads, and seismic loads.
In the first instance excessive loads generate a general foundation failure whereas seismic loads,
which by nature vary in time, may induce only permanent irrecoverable displacements.
Failure can therefore no longer be defined as a situation in which the safety factor becomes less than
unity; it must rather be defined with reference to excessive permanent displacements which impede
the proper functioning of the structure.

6.9.1 General requirements


One key requirement, stated in clause 5.1(1) of EN1998-5, is that the earthquake-induced permanent
deformations in the foundation remain small. Since foundations are placed underground and it is
difficult to inspect and repair them, excursions of the soil-foundation system into the plastic
deformation range are to be avoided even in severe earthquakes (with design accelerations as high as
0.3 – 0.4g on type A ground).

98
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

In order to restrain the horizontal displacements of the foundations, so that their response remains
essentially elastic with small residual displacement, allowable total horizontal displacement may be
assumed as a design target. The Japans standard for the design of foundation recommends 1% of the
width of the foundation as the allowable horizontal displacement, limited to 50 mm for large (> 5 m)
foundations. For foundation piles with a diameter not exceeding 150 cm the recommended limit is
15 mm.

6.9.2 Rules for conceptual design


Except line constructions such as bridges and pipelines, the mixed foundation type for example
structure founded partly on piles and partly with shallow foundations is not recommended (clause
5.2(1) of EN1998-5:2004). But if such a solution is unavoidable (which is quite usual in Norway
because of variation in depth to bedrock):
 It may be used seismic joints between units of structure with different foundation concept.
The seismic joint makes the units dynamically independent.
 Or, it shall be used if a specific study which demonstrate the adequacy of such a mixed
foundation solution.

Inclined piles are not recommended to transmit the lateral loads to the soil as entailed in clause
5.4.2(5) of EN1998-5. If so an appropriate analysis should be performed to demonstrate that the pile
system can safely carry both the axial and bending loads.

6.9.3 Transferring the structural loads to foundation – soil system


Design horizontal force shall transmit to the ground by means of:
 design horizontal shear resistance between the foundation base and the ground,

 design shear resistance between vertical sides of foundation and the ground,

 design resisting earth pressure in vertical front face of the foundation.

Normal forces and moments shall transmit to the ground by means of:
 design value of resisting vertical forces acting on the base of the foundation,

 horizontal shear resistance between sides of deep foundation and the ground,

 design value of vertical shear resistance between sides of the embedded and deep foundation and

the ground (boxes, piles, piers).

Concerning the transfer of the horizontal force to the ground, engineering judgment was used in
clause 5.3.2(3) of EN1998-5:2004 to allow a fraction of it to be resist by at most 30% of the fully
mobilized earth pressure on the front face of the foundation.
To gain a more quantitative insight, one may subdivide the total design horizontal force into
shear force, , acting at the bottom horizontal base of a footing of foundation slab, and into the
horizontal force, , acting on the vertical front faces of the foundation (where passive earth
pressure is mobilized), using the following expressions:

99
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

where is the allotment ratio of horizontal forces, is the coefficient of horizontal

Subgrade reaction (kN/m3) over the sides of the foundation, is the effective embedment depth
(meters), is the coefficient of shear Subgrade reaction (kN/m3) at the base of the foundation and
B is the width of the foundation (meters).
The coefficient may be computed as , where is typically obtained from an in situ

plate bearing test with a rigid disc of diameter , and is the loading area (m2) of the foundation
perpendicular to the load direction.
The coefficient of shear Subgrade reaction, , may be estimated as , where = 1/3 and
is the coefficient of vertical Subgrade reaction determined by an empirical expression similar to
that just given for .

100
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

6.10 Shallow Foundations


6.10.1 Soil stiffness
The response of the foundation as a system under earthquake loading is strongly affected by the
stiffness profile and depth of the foundation. These affect the fundamental frequency of the soil
layer, which in turn affects the amplification of bedrock motions to the surface and the foundation
system damping characteristics.
Three stiffness profiles are shown in Figure 6.6. The fundamental frequencies of those profiles are
given by:
Constant stiffness with depth:

Linearly increasing stiffness:

Parabolic stiffness with depth:


where is the shear wave velocity of the ground, and H is the depth.

Figure 6.6. Typical stiffness profiles of the soil.

Section 4.2.3 of EN1998-5 provides a table that gives average soil damping and reduction factors for
shear wave velocity and shear modulus, which can be used in the absence of specific
measurements/calculations.
The non-linear stress-strain response of soils result in different amplifications of the bedrock motion
through the soil column, depending on the magnitude of the earthquake motions (Idriss, 1990). The
greatest amplification of bedrock accelerations occurs at low peak acceleration levels. As the peak
acceleration level increases for larger earthquakes, so the amplification of the soil column decreases
and becomes less than unity at high bedrock acceleration levels. Observations from soft soil sites
suggest that the crossover from amplification to de-amplification occurs at bedrock accelerations of
about 0.3g to 0.5g (Mohammadioun and Pecker, 1984; Idriss, 1990; Suetomi and Yoshida, 1998).

6.10.2 Overview of behavior


The performance of shallow or spread foundations subject to seismic loading can be considered as
consisting of several modes as presented in Figure 6.8.
1. The long-term static loading will have produced some foundation displacement.

101
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

2. For relatively small seismic loadings most foundations will respond in an essentially linear
elastic manner.
3. As the loading increases towards the ultimate dynamic capacity, non-linear soil responses
become significant and the foundation response may be affected by partial uplift.
4. The ultimate capacity of the foundation will be significantly influenced by the dynamic
loadings imposed, with transient horizontal loads and moments acting to reduce the ultimate
vertical capacity. For transient loadings that exceed yield, permanent displacements may
occur.

Figure 6.7. Conceptual response of spread foundation to seismic loading.

Lateral loading may generate sliding with larger sliding displacements accumulating if the transient
horizontal loading is biased in one direction. Uplift and rocking behavior may result in permanent
rotations while bearing capacity failure will lead to settlement, translation and tilt.
In addition to the transient and permanent deformations that arise from loads transmitted through
the structure into the foundation, additional displacements may arise from ground movements
imposed on the foundation. In this class of behavior are settlements arising from densification of the
soil, the effects of liquefaction and lateral spreading.

6.10.3 Ultimate capacity of shallow foundations


EN1998-5 requires the ultimate seismic capacity of footings to be assessed for the both sliding and
bearing capacity ‘failure’.
Sliding
The friction resistance for footings on cohesionless deposits above the water table, , may be
calculated from the following expression:

where - design normal force on the horizontal based


 - interface friction angle

102
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

M - partial factor (1.25 for tan).


For cohesive soils the equivalent relationship is:

where A = plan area of foundation


cu - undrained strength
M - partial factor (1.4 for su).

Most foundations are embedded and derive additional resistance to sliding by mobilizing passive
resistance on their vertical faces. For some classes of foundation (e.g. bridge abutments) this
resistance provides a major contribution to their performance.
However, the mobilization of full passive resistance requires significant displacements, which may
amount to between 2% and 6% of the foundation’s embedment depth. Such displacements may
exceed the maximum allowable values for the structure and hence the foundation design may
incorporate only a proportion of the full passive resistance.
EN1998-5 requires that to ensure no failure by sliding on a horizontal base, the following expression
must be satisfied:

where: is the design lateral resistance from earth pressure, not exceeding 30 % of the full passive
resistance.

Bearing capacity
Static bearing capacity
Bearing capacity formulae for seismic loading are generally related to their static counterparts. For the
static case:

where q = ultimate vertical bearing pressure


c = cohesion
 = soil density
B = foundation width
= overburden pressure at foundation level
= bearing capacity factors
= shape factors
Closed form solutions exist for and but not for . Thus while the factors and have
widely accepted definitions, a considerable range of solutions have been proposed for based on
approximate numerical studies or on experimental results.
A selection of the suggested equations is presented in Table 6.5.

103
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

Table 6.5. Formulations for bearing capacity factors.

Inclined loading is incorporated into the bearing capacity equation either by incorporating inclination
factors into each term of equation or by direct modification of the bearing capacity factors. Thus:

where are inclination factors.


Various proposals for the inclination factors are shown in Table 6.6.
Moment acting on the foundation is treated by defining an effective foundation width, B´. The
horizontal and vertical loads are applied to the effective foundation. B´ is defined as follows:
and
where M = applied moment
H = horizontal loading (parallel to B)
V = vertical loading
A = plan area of foundation, BL

Table 6.6. Published relationships for inclination factors.

Seismic bearing capacity

104
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

Significant earthquake events substantially reduce the ultimate bearing capacity of spread footings
due principally to the following effects:
 The imposition of transient horizontal loads and moments arising from the inertia of the
supported structure.
 Inertial loading of the foundation material.
 Changes in the strength of foundation materials due to rapid cyclic loading.
Several solutions have recently been published for bearing capacity that take account of inertia effects
in the foundation material. The seismic bearing capacity may be expressed:

where q = vertical component of the ultimate bearing pressure


are seismic bearing capacity factors.
While this expression appears suitable for the evaluation of shallow foundation behavior on either
granular or cohesive soil, some caution is required. The rate of loading applied by seismic events is
sufficiently high to cause the response of a saturated granular stratum to be essentially undrained
beneath the footing.
Solutions have been presented as the ratios of the static to the dynamic bearing capacity factors for
cases where the shear transfer factor is 0, 1 or 2, see Figure 6.8.
and are substantially affected even in cases where the horizontal loading imposed by the
foundation remains at its static value.

Annex F of EC8-5 presents an alternative method for assessing bearing capacity of strip, shallow
foundations. The result is based on a long-term European research program, including field evidence,
analytical and numerical solutions and a few experimental results (Pecker and Salençon, 1991;
Dormieux and Pecker, 1995; 1997b; Pecker 1997).
The stability against seismic bearing failure of a shallow foundation may be checked with the
following inequality:

where for a footing of dimensions width B and length L:


, and

105
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

Figure 6.8. Seismic bearing capacity factors with horizontal acceleration and angle of internal
friction.

Table 6.7. Values of parameters.

106
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

where , and are the design action effects at the foundation level, and the rest of the
numerical parameters depend on the type of soil and are given in Tables 8.

Purely cohesive soils


The ultimate bearing capacity of the foundation under a vertical centered load is given by:

where is the undrained shear strength of the soil


s is the partial factor for the undrained shear strength.
The dimensionless soil inertia is given by:

where  is unit mass of the soil


is design ground acceleration on type A ground, given by
is reference peak ground acceleration
I is importance factor, depending on the building importance
S is soil factor.
The following constraints apply to the general bearing capacity expression:

Purely cohesionless soils


The ultimate bearing capacity of the foundation under a vertical centre load is given by:

where is vertical ground acceleration, given by


is bearing capacity factor, given by:

where is design shearing resistance angle given by:

The dimensionless soil inertia is given by:

The following constraints apply to the general bearing capacity expression:

where k is a coefficient from Table 8.

107
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

6.10.4 Foundation horizontal connections


Tie beams should be provided between all foundations, except for ground type A (rock), or on
ground type A and B (stiff soil) in areas of low seismicity. The tie beams should be designed to
withstand an axial force, considered in both tension and compression, equal to:
for ground type B
for ground type C
for ground type D
where is mean value of the design axial forces of the connected vertical elements

6.10.5 Design example on a shallow foundation

The large shallow foundation footing of the pier of a viaduct, located on the high speed train line is
to be verified against bearing capacity failure, the dimension of foundation are as follow:
• Plan dimensions: B = 11.4 m, L = 12.4 m
• Elevation of foundation base: +186.85 m
• Elevation of ground surface: +191.15
• Thickness of footing: 2.5 m

The design soil profile are:


Depth from ground Material Unit weight Angle of sharing Undrained shear
surface (m) (kN/m3) resistance  () strength (kPa)
0 – 14 Gravel 20.5 38 0
> 14 Gravel and sand 19.5 37 0

The most unfavorable combination of loads acting on the foundation, obtained from analysis of the
viaduct structure, is: , , .
The seismic action is determined knowing that the construction site lies in a low seismicity zone. The
bedrock design acceleration, . the ground at site is considered type B (deposit of dense to
very dense sand and gravel), for which a soil factor S = 1.25 is assumed. This gives a design ground
acceleration: .

Verification against bearing capacity failure


According to equation (F.6) in EC8-5, the ultimate bearing capacity, per unit length of the strip
foundation under vertical centered load is

where
= vertical ground acceleration =
= the bearing capacity factor, given by

108
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

where the design value of the angle of shearing resistance is given by ,


B = 12.4 m and .
Substituting the previous value gives

and, taking into account the actual length of the foundation and using the smaller value, yields the total
bearing capacity:

For medium-dense to dense sands, Table F.2 of Annex F gives . hence, substituting in equation
(F.2) of EC8-5 yields

For a purely cohesionless soil, the dimensionless inertia force is given by equation (F.7) of EC8-5

And the value of satisfies the required condition (F.8) of Annex F:

Furthermore

Substituting all the previous values and the appropriate numerical parameters into equation (F.1) now yields

Since the left-hand side equals -0.66, the inequality is satisfied, and safety against bearing capacity is
verified with a wide margin.

109
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

6.11 Pile Foundations


6.11.1 Examples of pile foundation failures following earthquake loading
The most important cause of pile damage and failure during the past earthquake was caused by loss
of lateral soil support that might occur for
1) liquefaction of granular soils under water table, or
2) strain softening of cohesive soils near the pile head.
The pile may experience a loss of bearing capacity and consequently loss of structural capacity of the
pile. When combined with large structural inertial loads and rocking mode induced by superstructure
inertia forces the pile can develop excessive displacements, bending strains concentrated near the pile
head or settlements, punching failure and tensile pull-out failure.
Batter piles are designed to accommodate large lateral loads, but they often attract forces that the pile
heads and/or pile cap cannot sustain.
In non-liquefiable soils, the piles may be also subject to damaging bending strains at interface
between soil layers of strong impedance contrast. The contrast may be provided by soft and stiff
layers, or by soil layers that undergo liquefaction or strain softening under earthquake loading.
Reinforced concrete has the disadvantage, compared to structural steel, of loosing progressively its
strength and stiffness upon cyclic loading. After several strain reversals bending and shear cracks may
intercross each other leading to an extensive disorganization of the material.
Pile failures can cause either collapse of the superstructure or excessive settlements and rotations.
More recently the building at Kandla Port suffered a rotation of about 11 from the vertical
following the Bhuj earthquake of 2001 as shown in Figure 6.9. The pile foundations supporting this
building have suffered differential settlement.
Tokimatsu et al (1997) describe the failure of a three-storey building supported on pile foundations
during the 1995 Kobe earthquake as shown in Figure 6.10.

110
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

Figure 6.9. Rotation of a tall building on pile Figure 6.10. Failure of piles in a three-story building
foundation during the Bhuj earthquake. in 1995 Kobe earthquake.

6.11.2 Eurocode 8 provisions


The normal static design of pile foundations must be carried out under the provisions of EN1997-1
(clause 5.4.2(2) of EN1998-5). In addition, EN1998-5 recommends a careful assessment of
construction site soil to determine the nature of supporting ground, slope instability, high
densification susceptibility and the liquefaction potential (clause 4.1.1(1) of EN1998-5).
It is also suggested that careful consideration of any additional loading on the piles and pile caps that
may arise due to the lateral spreading of the soil, particularly in the presence of a non-liquefiable soil
strata overlying a liquefiable layer.
In addition, where liquefaction is anticipated it is suggested that the strength of the liquefied soil
must be ignored (clause 5.4.2(4) of EN1998-5).

6.11.3 Flexibility of piles


General
Summaries of the methods used to assess the responses of piles and pile groups to seismic loading
are provided by Gazetas (1984), Novak (1991) and Pender (1993).
Numerical studies indicate that the response of a pile shaft under seismic loading can be considered
in three zones:
1. The near surface zone. This zone extends to approximately eight pile diameters beneath the
soil surface and is dominated by inertial loading effects.
2. An intermediate zone. This zone exists between the near surface and deep zones and is
influenced by both inertial and kinematic effects.
3. The deep zone. This zone is below 12 to 15 pile diameters from the surface and is dominated
by kinematic effects.

111
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

Pile flexibility
As an alternative to the method suggested above, the flexibility of the pile can be determined using
the following procedure. The elastic length of pile can be determined using elastic length of pile, T:

where is the flexural stiffness of the pile,


k is “gradient of the soil modulus”,
which in cohesionless soils, k takes values between about 2000 kN/m3 for loose saturated conditions
and about 20,000 kN/m3 for dense conditions above water table.
For normally consolidated cohesive materials, k may vary between about 200 kN/m3 and
2000 kN/m3.
For a soil in which the elastic modulus  0 at the pile head, i.e. when the Young modulus can be
expressed in the form , the elastic length T van be written as

A non-dimensional length is defined as:

Where is the actual length of the pile. Based on these definitions, the following classification is
introduced:
 if > 5, the pile is flexible, i.e. its behavior is not affected by the length, and collapse is
always caused by a flexural failure, with formation of a plastic hinge.
 if 5 > > 2. 5, the pile is semi-flexible, i.e. its behavior is affected by the length, and
collapse may be caused either by flexural failure, or by attainment of the ultimate resistance of
the soil.
 if < 2.5, the pile is rigid, i.e. it behaves like a pier – in this case, the flexural
deformation can be neglected with respect to rigid rotation, and the collapse always occurs
because the ultimate resistance of the soil is attained.

Piles that are classified as flexible will ‘move’ with the surrounding soil and therefore would attract
the inertial shear load imposed by the superstructure during earthquake loading.
Rigid piles, on the other hand, will attract significant soil load, as the piles stay in position and the soil
would exert passive pressures on either side of the piles in alternative load cycles. This additional
lateral load applied by the soil must be considered in the pile design.
The soil-pile interaction for rigid piles (piers) may be analyzed, making reference to limiting
equilibrium solutions. However, finite-element analysis should be employed when pier head
displacements and rotations need to be accurately evaluated.

6.11.4 Seismic actions on the piles


Clause 5.4.2.(1) of EN1998-5:2004, states that piles and pers shall resist two types of action effects:

112
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

• Inertia forces from the structure combined with static loads


• Kinematic forces arising from deformation of the surrounding soil due to passage of seismic
waves.
The effects of the inertia forces acting at the pile head are strongly reduced at depths greater than 3.0
– 3.5 times T, so that, in practice, the pile response depends to extent on the soil properties within
the same depth range.
Therefore, elastic pile-soil interaction remains a viable assumption as long as the shallow soil layers
near the pile head do not suffer rupture or significant plastic deformation.
Otherwise, the elastic theory is no longer applicable, and one has to resort to a fully nonlinear
approach, such as so-called p-y curves.
The need to take the nonlinearity of the soil-pile interaction into account is also strongly influenced
by the pile diameter: when small-diameter (e.g. 0.2 – 0.3 m) or micro-piles, are used, elastic theory is
frequently not applicable.

6.11.5 Pile fixity conditions


At the ground surface a pile may be free to rotate or connected to some structure so that is
constrained. These are so-called free head and fixed head piles. The idea is illustrated in Figure 6.11.
For the fixed head pile a moment is required to enforce the compatibility requirement that the
rotation of the pile head is equal to that of the pile cap.
As can be seen from the diagram the fixing moment has the opposite sense to a moment caused by
the application of the horizontal shear above the soil surface, thus it carries a negative sign.
It is obvious that the pile head fixity condition has effect on both translational and mostly on
rotational stiffnesses of the pile.

Figure 6.11. Free and fixed head piles.

6.11.6 Kinematic and inertial loading


For many classes of structure the predominant static loading on piled foundations is vertical
compressive loading. Earthquake loading will impose requirements on the piles to resist significant
lateral loads and moments with the further possibility of piles being required to carry tensile loads.
The deformation of piles may be substantially affected by the permanent deformations of the
ground.

113
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

Earthquake loading differs from other forms of environmental and machinery induced cyclic loading
because the in-ground motions produce pile loadings in addition to the pile loadings derived from
the motion of the supported structure. The in-ground motion generates ‘kinematic interaction’
between the piles and the soil while the loading imposed by the structure generates ‘inertial
interaction’ as shown in Figure 6.12.

Figure 6.12. Kinematic and inertia interaction.

EN1998-5, clause 5.4.2(6) notes that bending moments due to kinematic interaction only need to be
considered when all the following conditions apply:
• The ground profile is of type D, S1 or S2 and contains consecutive layers of sharply differing
stiffness.
• The zone is of moderate or high seismicity (i.e. .S exceeds 0.1g),
• The structure is of importance class III or IV.

6.11.7 Ultimate lateral soil resistance


The determination of the lateral soil resistance for laterally loaded piles is a complex problem of the
ULS of an elastic plastic medium, for which rigorous closed-form solutions are not available.
As a first approximation:
• for piles embedded in cohesionless soils the simple expression has been proposed:

where a is an adjustment factor taking values between 3 and 4.


• for piles in cohesive materials under undrained static loading condition:
for

for
The value for is determined for . It maybe recall that, for normally consolidated
materials, is typically a linear function of depth.

114
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

Figure 6.13. Ultimate pressure distribution against a laterally loaded pile in cohesive soil, a) free head and 2)
fixed head pile.

6.11.8 Kinematic response


It is convenient to analyze the kinematic response of the pile or pile group separately from the
inertial response. The kinematic response at depth may be used to assess the structural requirement
of the pile in the intermediate and deep zones. The kinematic response of the pile head is an input
into the inertial response analysis.
In the deep zone the presence of piles has little effect on the ground motion or natural frequency of
the stratum. The pile and soil motions are likely to be practically coincident for frequencies up to at
least 1.5 times the natural frequency, , of the ground. This observation is of practical significance as
the deflected shape of the pile can be obtained from a 1D equivalent linear shear wave propagation
analysis. Having obtained the deflected shape of the pile, its bending moments and shear forces may
readily be determined.
EN1998-5 requires piles to remain elastic, though under certain conditions they are allowed to
develop plastic hinges at their heads. The regions of plastic hinging should be designed according to
EN1998-1, clause 5.8.4.

6.11.9 Inertial response


The inertial response analysis uses the dynamic response obtained from the kinematic interaction
study to assess the seismic displacements and rotations of the pile head or of the structure. The
forces driving the pile head are derived from the mass and stiffness of the structure.
Typically the structure may be simplified to a single degree of freedom system while the piled
foundation is considered to have translational and rotational degrees of freedom.

6.11.10 Combination rules


Since the response of the soil and structure will be at different natural frequencies, the combination
rules given in clause 4.3.3.5 of EN1998-1 can be used to calculate the cumulative effect of kinematic
and inertial interaction.

115
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

6.11.11 Norwegian ”Peleveiledning 2005”


Chapter 11 in Norwegian guidance for pile foundation (“peleveiledning 2005”) discusses the seismic
resistance design of pile groups. The procedure presented in this guidance following the same
approach presented here and it is valid also with respect to Eurocode 8 as presented in Figure 6.14.

Figure 6.14. Flowchart for seismic analysis and design of pile foundations (Peleveiledning, 2005).

116
EC8 Guide Book Chapter 6 – Geotechnical aspects and foundations

6.11.12 Construction Detailing


Although the safety of a constructed facility does not rely only upon a blind application of seismic
codes and standards which are used for its design and construction, those documents help
significantly to minimize the most commonly encountered causes of deficiencies and failures.
Because all phenomena described previously cannot be analyzed with the necessary mathematical
rigor and are not often relevant to even sophisticated calculations, construction detailing must always
be enforced in seismic design of foundations.
Many of these detailing practices, which are found in the most recent codes, are little more than
common sense. However, they represent the most common mistakes made in design by non-
experienced designers:
• Liquefiable deposits and unstable slopes must always be treated before construction. Even if
a piled foundation can survive the cyclic deformations of a liquefied deposit, the quasi static
displacements imposed by the post-earthquake ground flow are an order of magnitude larger
and cause distress of the foundation as evidenced in the Kobe earthquake.
• The foundation system under a building must be as homogenous as possible unless
construction joints are provided in the structure.
• In particular, for individual footings, the situation where some of them rest on a man-made
fill and some on in-situ soils must always be avoided.
• It is highly desirable that the foundations respect the symmetries of the building.
• The choice of the foundation system must always account for possible secondary effects such
as settlements in medium-dense or loose dry sands, the post-earthquake consolidation
settlements of clay layers, the settlements induced by the post-earthquake dissipation of pore
pressures in a non liquefiable sand deposit.
• Raft foundations or end bearing piles are to be preferred whenever the anticipated magnitude
of the settlements is high or when they can be highly variable across the building.
• Individual footings must always be linked with tie beams at the foundation level. These
longitudinal beams must be designed to withstand the differential settlements between the
footings.
• Piles must be reinforced along their whole length, even if calculations do not require
reinforcement. Special care must be given to the connections with the raft or to soil layers
interfaces when two layers in contact present marked differences in stiffness. For instance,
the connection with the raft can be detailed to allow for a plastic ductile hinge as allowed in
Eurocode 8.
• Inclined piles must preferably be avoided: they can be subjected to parasitic bending stresses
due to soil densification following an earthquake, they induce large forces onto pile cap and,
if their arrangement is not symmetric, permanent rotations may develop due to different
stiffness of the pile group in each direction of loading.

117

You might also like