Вы находитесь на странице: 1из 63

1

1
2
2

SOIL-FOUNDATION-
STRUCTURE-INTERACTION
(SFSI)



12.1 INTRODUCTION

In chapter 11 we considered soil-structure interaction, but mainly from the
classical consideration of the influence of elastic behavior of the soil beneath
the foundation. We found, Example 11.1 a water tank supported on a shallow
foundation, that when truly elastic behavior the soil supporting the foundation
is considered the contribution from soil-structure interaction is very small. We
then extended this to represent earthquake induced nonlinear soil behavior by
reducing the soil modulus in the manner recommended in Tables 11.1 and
11.2. Even then we found that a very substantial reduction in soil modulus was
required to give significant soil-structure interaction effects. Later in the
chapter we the made the model more realistic, Example 11.2, by including the
mass of the foundation block supporting the water tower. This additional mass
meant that the number of degrees of freedom was increased from 1 to 3 but
the size of the foundation required to satisfy ultimate limit state conditions
could be reduced because of the stabilizing effect of the additional weight.
Once again we concluded that soil-structure interaction effects were not of
great significance. For this model there was more than one natural period (or
frequency) for the system but the equivalent system period is likely to be of
most significance during an earthquake. An additional limitation of all the work
in Chapter 11 is the implicit assumption that there is always full contact
between the underside of the foundation and the supporting soil and that there
is always a reserve of bearing strength in hand. The purpose of this chapter is
to present four separate strands of information. First we will consider data
from field testing of shallow foundations at a site in Albany. Nonlinear
foundation behaviour was an important part of these tests. Second, we will
Earthquake Resistant Design of Foundations


10

introduce the macro-element idea which provides a means of allowing for
bearing strength failure of the soil beneath the foundation (this imposes limits
on possible actions that the foundation can sustain). Third, we will compare
the insight so obtained with information gained from modeling the foundation
as a bed of nonlinear springs that can detach and reattach as the foundation
responds to the earthquake excitation. Finally, an example is given of the
design of a shallow foundation for an unreinforced masonry building following
a displacement approach.


12.2 SHALLOW FOUNDATION NONLINEAR PUSHOVER RESPONSE

Field experiments have been conducted at a site in Auckland with both shallow
and deep foundations subject to cyclic loading. More details are given by Algie
et al (2010) and M.Sadon et al (2010). The first batch of tests used an eccentric
mass shaking machine to excite the foundations with sinusoidal oscillations at a
range of frequencies. Although successful there are limitations to this approach
for the following reasons. First a given level of excitation force cannot be
obtained until the shaker frequency has been increased from zero to the
frequency required to generate the required force. Second the response of the
system is measured under steady state excitation at a fixed frequency. In this
way what is obtained from the use of a shaking machine is not representative
of what happens during earthquake excitation.

An alternative, described here, is the use of snap-back testing. This test is
simpler than using an eccentric mass shaking machine. It gives the response of
the system to one impulsive excitation instead of continuous excitation; it is
more representative of what occurs during an earthquake. An added bonus is
the static load-deflection curve obtained during pull-back phase of the test. The
initial pull-back can generate a force of comparable magnitude to the
maximum force that can be produced by the shaking machine used.

Results are presented for the nonlinear stiffness and damping of shallow
foundations from snap-back testing. Tests were done at a site with Auckland
residual clay. A sequence of snaps from different initial loads shows how the
nonlinear behaviour of the foundation develops as the applied load increases.

The site used for the tests, in Albany in the northern part of Auckland, consists
of a profile of stiff cohesive soil formed by in situ weathering from tertiary age
sandstone and siltstone (it is thus a residual soil profile). There were eight
shallow foundations, which support the ends of the steel frame shown in
Figure 12.1a; these are reinforced concrete 2.0 m in length and 0.4 m square.
The steel frame structure is 2 m wide, 3.5 m high and 6 m long. Steel kentledge
is strapped to the top of the frame to provide the required vertical foundation
load. The soil profile was investigated with 21 CPT tests between the surface
and depth of 5 or 8 m; in some of these the shear wave velocity of the soil was
measured. The s
u
values obtained from the CPT q
c
values are reasonably
consistent with depth at about 100 kPa. Hand shear vane testing also done.
The shear wave velocity measurements from the seismic cone penetration tests
were supplemented with WAK tests (Briaud and Lepart 1990) and SASW tests
Chapter 12: Soil-Foundation-Structure-Interaction

11

(Stokoe et al 1994, Stokoe et al 2004). All of these indicated a reasonably
consistent shear wave velocity for the materials at the site equivalent to a small
strain shear modulus for the soil of about 40 MPa.




(a) Shallow foundation mounted demountable steel structure




(b) Snapbacktestarrangement,showingthereactioncrane,chainstoapply
thepullbackforce,thesteelframewithsteelkentledgeinplace.

Figure 12.1 Shallow foundation configuration and set-up for snap-back
testing
Earthquake Resistant Design of Foundations


12


0
1
2
3
4
5
6
7
8
9
10
0 2 4 6 8 10
Cone Resistance, q
c
(MPa)
D
e
p
t
h

(
m
)
0.00 0.05 0.10 0.15 0.20
Sleeve friction, f
c
(MPa)
qc
fc
0
1
2
3
4
5
6
7
8
9
10
11
0 100 200 300 400 500
Undrained Shear Strength, s
u
(kPa)
D
e
p
t
h

(
m
)
GL: 0.0m
Figure 12.2 Representative CPT data from the Albany site.

The CPT profiles in Figure 12.2 indicate a reasonably consistent q
c
values for
about the upper 5 m of the soil profile. The water table is noted to be at a
depth of 5 m, this reflects the time of the year at which the CPT tests were
done (June 2009) but since the soil is cohesive it is saturated to near the ground
surface the whole year. Many of the tests were done shortly after rain so that
free water was lying in any pockets on the ground surface (cf Figure 12.1b).

The equipment required for snap-back testing is simple. Instrumentation and
data logging equipment are, of course, the same as needed for the testing with
the eccentric mass shaking machine. To apply the snap-back force a hydraulic
jack, load cell and quick release mechanism are needed. In addition a reaction
force has to be mobilized against the cable applying the snap-back force; in our
case the crane located at the site provided this. The quick release device used
was obtained from a yacht chandler and is one of a range used for releasing
spinnakers quickly. The device obtained, the largest available, had a working
force capacity up to 100 kN. The set-up for the shallow foundation tests is
shown in Figure 12.1b.

In Figure 12.3a are shown all the Test 7 static moment-rotation curves
obtained during the application of the pull-back forces. Figure 12.3b does the
same for the Test 9 pull-backs. It is apparent that there is considerable
nonlinearity in the moment-rotation curves and also that the stiffness is
reduced from one snap-back to the next, in particular for those tests following
the snap-back which applies the largest moment to the system.


Chapter 12: Soil-Foundation-Structure-Interaction

13

0 5 10 15 20 25 30
0
20
40
60
80
100
Rotation (millirads)
M
o
m
e
n
t

(
k
N
m
)


Snap 1
Snap 2
Snap 3
Snap 4
Snap 5
Snap 6
Snap 7
Snap 8
Snap 9

0 5 10 15 20
0
20
40
60
80
100
Rotation (millirads)
M
o
m
e
n
t

(
k
N
m
)


Snap 1
Snap 2
Snap 3
Snap 4
Snap 5
Snap 6
Snap 7
Snap 8
Snap 9

Figure 12.3 Moment-rotation curves obtained during the pull-back parts of
the snap-back tests; (a) Test 7 and (b) Test 9.

a
b
Earthquake Resistant Design of Foundations


14

0 0.05 0.1 0.15 0.2 0.25 0.3
0
0.2
0.4
0.6
0.8
1
1.2
Normalised foundation rotation
N
o
r
m
a
l
i
s
e
d
m
o
m
e
n
t
c
a
p
a
c
i
t
y Normalised moment capacity at applied vertical load
Best fit curve through measured moment-rotation data
Measured moment-rotation data
a
0 0.05 0.1 0.15 0.2 0.25 0.3
0
0.2
0.4
0.6
0.8
1
1.2
Normalised foundation rotation
N
o
r
m
a
i
l
s
e
d
s
e
c
a
n
t
r
o
t
a
t
i
o
n
a
l
s
t
i
f
f
n
e
s
s
Foundation rotational stiffness based on G
max
b
10
-5
10
-4
0.001 0.01 0.1 1
0
0.2
0.4
0.6
0.8
1
1.2
Normalised foundation rotation (log 10 scale)
N
o
r
m
a
l
i
s
e
d
s
e
c
a
n
t
r
o
t
a
t
i
o
n
a
l
s
t
i
f
f
n
e
s
s
Foundation rotational stiffness based on G
max
Measured data
c
Rotation
M
o
m
e
n
t
1
K
_secant


Figure 12.4 Curve-fitted moment-rotation relations matched to the recorded
data for the first three pull-backs of Test 7 and Test 9. (a) moment-rotation
data, (b) and (c) secant modulus against rotation.

Chapter 12: Soil-Foundation-Structure-Interaction

15

12.2.1 Interpretation of the field test data

In Figure 12.4 the data presented in Figure 12.3 for the initial three pull-backs
of Test 7 and Test 9 are processed further. Each of the graphs in Figure 12.4
has the data plotted in a normalised form.

In Figure 12.4a a best-fit curve based on the stiffest test results of Test 7 and 9
is shown. The reason that the stiffest response is shown is because after the
first few snap-back tests there is degradation of the foundation stiffness,
presumably because of the irrecoverable deformation of the ground beneath
the foundation. Also shown in this figure is the estimated moment capacity of
the foundation given the applied vertical load. Figure 12.4b has plotted the
normalised rotational stiffness of the foundation. The rotational stiffness is
normalised with respect to the elastic rotational stiffness obtained from the
small strain shear modulus of the soil (the value used to obtain the maximum
value was 30 MPa). The horizontal axis on this diagram is the normalised
foundation rotation; the normalisation parameter being the rotation at which
the structure would topple (for the structure tested this is 0.3 radians).
However, Figure 12.4b is not particularly useful as the rotational stiffness
decreases so rapidly at small foundation rotations. A more useful view of this
data is given in Figure 12.4c, which is simply Figure 12.4b replotted with a
logarithmic scale for the rotation axis. Also noted in Figure 12.4c is the range
of rotations over which data was recorded. The graph shows very clearly that
the test data covers the range of rotations over which there is a rapid reduction
in rotational stiffness.


12.3 MEASURED NONLINEAR DYNAMIC EXCITATION OF SHALLOW
FOUNDATIONS

In this section the above work on the pull-back phase of snap-back tests is
extended by presenting results of dynamic loading of shallow foundations
the loading having been generated by the use of an eccentric mass shaking
machine and by snap-back testing. As explained above the purpose of these
field tests is to obtain nonlinear moment-rotation data for shallow foundations.

Figures 12.5 (a) and (b) show the steady state response of one of the shallow
foundation pairs at gradually increasing frequencies generated with an eccentric
mass shaker. The diagrams show that many cycles are required prior to
applying large amplitude cyclic moment. Through the origin a line is drawn
which defines the rotational stiffness of the shallow foundation based on the
small strain shear modulus of the soil (~ 40 MPa) with full contact between the
soil and the underside of the foundation. Clearly all the cycles plotted indicate
that the operational soil modulus is less than 40 MPa. Also plotted in the
diagram are bounds on the moment capacity defined by the bearing strength of
the soil (Algie et al 2010) from which one can see that the maximum applied
moment induces bearing failure in the soil beneath the foundation.

Figures 12.6 and 12.7 show snap-back results for some of the shallow
foundation tests. From Figure 12.6 the damping during the foundation rocking
Earthquake Resistant Design of Foundations


16

is evident and Figure 12.7 shows that there is energy dissipation during the
foundation rocking because of the hysteresis loops.

Figure 12.6 plots the response of snap-backs 1, 6 and 9 of Test 7. Note that the
maximum rotation for snap-back 6 is about an order of magnitude greater than
those for the other two. The response for snap-back 6, the middle plot in
Figure 12.6, shows a small amount of permanent rotation when the dynamic
response comes to an end. The damping values, determined by logarithmic
decrement during the first half cycles, are 42% for snap-back 1, 32% for snap-
back 6 and 34% for snap-back 9. These damping values are large in relation to
values usually applied in structural and foundation design. The decrease in
damping value going from snap-back 1 through to 9 is a consequence of the
accumulation of permanent deformation in the soil beneath the foundation, so
the effective length of the foundation will be decreasing gradually.

Figure 12.7 presents moment-rotation information calculated from data
recorded during two of the snap-back tests. Also included in the diagrams are
the data from the initial pull-back parts of the tests. The moments were
calculated from data obtained from strain gauges attached to the legs of the
steel frame structure.

Figure 12.8 has all the damping values obtained from snap-back testing of the
steel frame on the shallow foundations. The most important feature of this
diagram is the large values obtained for the damping parameter. The next most
significant feature of the diagram is the amount of scatter present. A factor
contributing to this will be the accumulation of permanent deformation beneath
the shallow foundations as the number of snap-back tests on a particular
foundation increases. This scatter will be related to the fact the moment-rotation
plots in Figure 12.3, coming from the same foundation, do not all fall along the
same curve. Centrifuge tests on rocking foundations have also shown no apparent
relationship between the damping value and the amplitude of rotation, Gajan and
Kutter (2008).

In a famous paper, Housner (1963), a relationship was given between the period
of a rigid block rocking on a rigid surface and the initial angle of rotation. The
early part of Housners curve is plotted in Figure 12.9 (the parameter o on the
horizontal axis is the angle of tilt at which the block would fall over, for the
structure shown in Figure 12.1 this angle is about 0.3 radians). Also plotted in
Figure 12.9 are the half periods measured during the rocking response in the snap-
back tests which are seen to fall along the Housner curve.


Chapter 12: Soil-Foundation-Structure-Interaction

17

-10 -8 -6 -4 -2 0 2 4 6 8 10
-60
-40
-20
0
20
40
60
Rotation (millirads)
M
o
m
e
n
t

(
k
N
m
)
Bearing strength bound
Bearing strength bound
Rotational stiffness based on
small strain modulus of the soil



-4 -3 -2 -1 0 1 2 3 4
-120
-80
-40
0
40
80
120
Rotation (millirads)
M
o
m
e
n
t

(
k
N
m
)
Bearing strength bound
Bearing strength bound
Rotational stiffness based on
small strain modulus of the soil


Figure 12.5 Steady-state forced vibration response of the shallow foundation
structure; (a) test ?, (b) test 4.






a
b
Earthquake Resistant Design of Foundations


18

0 1 2 3 4 5
-4
-3
-2
-1
0
1
Time (secs)
R
o
t
a
t
i
o
n

(
m
i
l
i
r
a
d
s
)
0 1 2 3 4 5
-30
-20
-10
0
10
Time (secs)
R
o
t
a
t
i
o
n

(
m
i
l
i
r
a
d
s
)
0 1 2 3 4 5
-4
-3
-2
-1
0
1
2
Time (secs)
R
o
t
a
t
i
o
n

(
m
i
l
i
r
a
d
s
)


Figure 12.6 Time histories for snap-backs 1, 6 and 9 of Test 7 on the shallow
foundation structure. (First half cycle damping values: snap-back 1 =42%,
snap-back 6 =32%, snap-back 9 =34%)

-15 -10 -5 0 5 10 15 20 25
-80
-40
0
40
80
120
Rotation (millirads)
M
o
m
e
n
t

(
k
N
m
)
-15 -10 -5 0 5 10 15 20 25
-80
-40
0
40
80
120
Rotation (millirads)
M
o
m
e
n
t

(
k
N
m
)


Figure 12.7 Moment-rotation loops for two Test 7 snap-back responses of the
shallow foundation structure (left: snap-back 4, right: snap-back 5)
Chapter 12: Soil-Foundation-Structure-Interaction

19

10
-1
10
0
10
1
10
2
0
10
20
30
40
50
60
Half Amplitude Rotation (millirads)
D
a
m
p
i
n
g

R
a
t
i
o

(
%
)


Test 5 east
Test 6 east
Test 6 west
Test 7 east
Test 7 west
Test 9 east
Test 9 west

Figure 12.8 Damping values measured during the snap-backs of the shallow
foundation structure

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
Normalised Rotation - u
0
/o
H
a
l
f

P
e
r
i
o
d

(
s
e
c
s
)


Housner's equation
Test 6
Test 7
Test 9


Figure 12.9 Housners initial rotation - rocking period curve (o is the angle at
which the structure would topple, about 0.3 radians) plotted with the
measured half periods during the snap-back responses







Earthquake Resistant Design of Foundations


20

12.4 SHALLOW FOUNDATION MACRO-ELEMENT

In recent years several researchers have pointed out that the LRFD requirement,
that no more than half the bearing strength should be mobilised during
earthquake shaking, is very conservative. Since the earthquake applies a high
frequency dynamic load it is proposed that the consequences of brief instances
when the LRFD requirement is infringed and even brief instances where the
bearing strength is fully mobilized are unlikely to lead to problems. Paolucci
(1997), Pecker et al (2011), Gazetas et al (2007) and (2011), Deng et al (2009),
Toh and Pender (2009), Sullivan et al (2010), Pender et al (2009), and others, all
make the point that if gains are to be realised from soil-structure interaction then
nonlinear soil behaviour must be mobilised. An early contribution to this line of
thinking is the paper of Taylor et al (1981). As explained in Chapter 1 when there
is truly nonlinear behavior of the soil beneath the foundation and loss and re-
establishment of contact over part of the foundation-soil interface, the term Soil-
Foundation-Structure-Interaction (SFSI) has come to replace soil-structure-
interaction (SSI) which has the connotation of elastic soil behaviour.

A macro-element is a single computational entity that represents the behaviour
of a shallow foundation and the adjacent soil (Figure 12.10). The approach is
the antithesis of the finite element technique which requires discretisation of a
substantial volume of soil beneath the foundation and is therefore not a
practical design tool. From the macro-element, only the response at a
representative point (the foundation centre) is obtained, but this gives a useful
approximation to the foundation response. In this way the macro-element is a
practical option for modelling soil-foundation-structure interaction.

The model used herein was developed by Toh (2008). It is a version of the
macro-element model developed by Paolucci (1997). Although the model is
simple compared to other shallow foundation macro-elements (for example,
Cremer et al, 2001, and Chatzigogos et al, 2007), it provides results that
compare well to those from experimental results and yielding spring models
(Pender et al, 2009).

We saw in chapters 7 and 8 bearing strength surfaces define the combinations
of vertical load, horizontal shear, and moment that cause bearing failure
beneath a shallow foundation.


Chapter 12: Soil-Foundation-Structure-Interaction

21


S
i
n
g
l
e
d
e
g
r
e
e
o
f
f
r
e
e
d
o
m
s
u
p
e
r
s
t
r
u
c
t
u
r
e
m
o
d
e
l
R
i
g
i
d
f
o
u
n
d
a
t
i
o
n
b
l
o
c
k
O
u
t
e
r
z
o
n
e
w
i
t
h
e
l
a
s
t
i
c
s
o
i
l
b
e
h
a
v
i
o
u
r
o
n
l
y
M
a
c
r
o
e
l
e
m
e
n
t
Nonlinear soil

a b
Figure 12.10 Macro-element model and the Eurocode 8 undrained bearing
strength surface.

The macro-element model used in this chapter is based on the Eurocode 8
bearing strength surface for cohesive soils (Figure 12.10b). The surface acts as
a yield surface in that, when the actions locus lies inside the surface, behaviour
is considered to be elastic, while, when the actions locus reaches the surface,
bearing failure occurs and behaviour is considered to be perfectly plastic. In
this manner, the macro-element is able to capture nonlinear foundation
response.

12.4.1 Example design scenario

To demonstrate the relative performance of the three alternative design
approaches considered, the macro-element was used in an example design
scenario to model a bridge pier founded on clay with an undrained shear
strength of 100 kPa. The static factor of safety was varied by changing the size
of the square, surface foundation. Three foundation approaches are compared:
standard LRFD with strength reduction factor of 0.5, a variation in which
bearing strength failure is just prevented under the seismic actions, and the
third is to proportion the foundation to achieve a bearing strength factor of
safety under static loading of 3 and allow bearing failure during earthquake
loading (this is referred to as a yielding foundation below).

The macro-element used requires minimal input parameters, all of which are
easily obtained. Table 12.1 presents the structural parameters used in the
model. The structural parameters were selected to represent a hypothetical
bridge pier that behaves as a single-degree-of-freedom oscillator in the
transverse direction. The foundation stiffness and damping values are varied
depending on the size of the foundation, and were calculated using formulae
presented by Gazetas (1991).



Earthquake Resistant Design of Foundations


22

Table 12.1 Structural parameters for example design scenario.
___________________________________________________
Height of superstructure centre of gravity
above centre of foundation 15 m
Effective mass of superstructure 2 10
6
kg
Mass of foundation 5 10
5
kg
Rotational inertia of superstructure
and foundation 5.15 10
6
kg.m
2

Horizontal stiffness of superstructure 4.94 10
5
kN/m
Horizontal damping of superstructure 6280 kN.s/m

___________________________________________________
0
0.2
0.4
0.6
0.8
1
1.2
0 1 2 3 4
Period (seconds)
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
El Centro, 1940 (scaled)
NZS1170.5 design spectrum
(Christchurch, shallow soil)

Figure 12.11 Response spectrum of earthquake record overlaid on
NZS1170.5 design spectrum.

An earthquake acceleration time history from the 1940 El Centro earthquake
(scaled up marginally) was used as the dynamic input to the model. The
response spectrum of this time history (Figure 12.11) closely represents the
design response spectrum at a shallow soil site for an Importance Level 4
structure (an essential facility or one with post-disaster function) in
Christchurch, New Zealand, as specified by NZS1170.5:2004. In a real design
scenario, more than one earthquake record would usually be adopted (for
example, under NZS1170.5:2004, three or more records must be used). For
clarity, only one record was used in this example scenario.

Behaviour under alternative design approaches
Figure 12.12 shows the path followed by the actions locus in [M, H] space (i.e.
within a section of the bearing strength surface at a constant vertical load V)
for all three design approaches considered. Overlaid are the outlines of the
different sized bearing strength surfaces of each design approach.

For this particular example scenario, a static factor of safety of three was
adopted for the yielding foundation design approach, although it is noted that
any foundation with a seismic factor of safety less than one will experience
some yielding. Figure 12.12 shows how, for the yielding foundation design
approach, occasional instances of bearing failure occur during the earthquake,
as the actions locus momentarily reaches the bearing strength surface.

It was found that a static factor of safety of eight was required to only just
prevent seismic bearing failure from occurring (seismic factor of safety of

Chapter 12: Soil-Foundation-Structure-Interaction

23


.
-300,000
0
300,000
-60,000 0 60,000
M (kNm)
H (kN)
16
8
3
Static FoS

Figure 12.12 Bearing strength surface and path followed by foundation
actions locus under various design approaches.

0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 2 4 6 8 10 12 14 16 18
Static factor of safety
S
y
s
t
e
m

n
a
t
u
r
a
l

p
e
r
i
o
d

(
s
)
Y
i
e
l
d
i
n
g

f
o
u
n
d
a
t
i
o
n

d
e
s
i
g
n

a
p
p
r
o
a
c
h
S
e
i
s
m
i
c

F
o
S

=

1
T
r
a
d
i
t
i
o
n
a
l

a
p
p
r
o
a
c
h
Fixed base structural period

Figure 12.13 Natural period of soil-foundation-structure system under various
design approaches.

unity). Figure 12.12 shows the path of the actions locus almost reaching, but
not touching, the bearing strength surface for a static factor of safety of eight.

Based on this, the static factor of safety required to satisfy traditional design
approach requirements is 16. Note that the factor of safety is defined in terms
of vertical action, not moment or shear, so the dimensions of the bearing
strength surface in the [M, H] plane do not double when the static factor of
safety is doubled. This is demonstrated in Figure 12.13, which shows that for a
static factor of safety of 16, the actions locus passes further than half of the
distance from the origin of the [M, H] plane to the bearing strength surface.

The definition of factor of safety in terms of vertical loads only can therefore
give a misleading impression of the reserve of bearing strength available under
lateral loading (when the actions locus moves mainly within the [M, H] plane).

Earthquake Resistant Design of Foundations


24

0
0.2
0.4
0.6
0.8
1
1.2
0 1 2 3 4
Period (seconds)
S
p
e
c
t
r
a
l

a
c
c
e
l
e
r
a
t
i
o
n

(
g
)
Static FoS = 3 (T
n
1.35 s)
Static FoS = 8 (T
n
0.79 s)
Static FoS = 16 (T
n
0.51 s)

Figure 12.14 Natural period of soil-foundation-structure system compared
with El Centro earthquake response spectrum.

Figure 12.14 shows how the natural period of the soil-foundation-structure
system varies with static factor of safety. The natural periods were obtained by
analysing response spectra. Overlaid on the plot is the fixed-base period of the
single-degree-of-freedom superstructure. As the static factor of safety
increases, the soil-foundation-structure system becomes stiffer, and the system
period decreases towards the fixed-base period of the structure. Under the
traditional design approach, the foundation is so stiff that the system period is
very close to the fixed-base period of the structure and the effect of soil-
foundation-structure interaction is minimized; as we found in Example 11.1.

The consequences of the different system natural periods obtained under the
three alternative design approaches depend on the characteristics of the design
earthquake spectrum. The system natural periods for the three design
approaches overlaid on the earthquake response spectrum for this design
example are shown in Figure 12.14.

For the particular earthquake acceleration time history used, the spectral
acceleration at the system natural period increases as the system natural period
decreases. Consequently, the foundation designed according to the traditional
approach experiences the highest spectral acceleration, whereas the yielding
foundation experiences the lowest spectral acceleration. This has important
implications for the foundation and structural actions, as discussed in the
following section.

Figure 12.15 presents plots of the structural action and foundation moment for
a range of static factors of safety. The factors of safety corresponding to the
three alternative design approaches are marked. Similar behaviour is observed
for foundation shear.

Chapter 12: Soil-Foundation-Structure-Interaction

25

0
5,000
10,000
15,000
M
a
x
i
m
u
m

s
t
r
u
c
t
u
r
a
l

a
c
t
i
o
n

(
k
N
)
0
50,000
100,000
150,000
200,000
250,000
0 2 4 6 8 10 12 14 16 18
Static factor of safety
M
a
x
i
m
u
m

f
o
u
n
d
a
t
i
o
n
m
o
m
e
n
t
t


(
k
N
m
)
Y
i
e
l
d
i
n
g

f
o
u
n
d
a
t
i
o
n

d
e
s
i
g
n

a
p
p
r
o
a
c
h
S
e
i
s
m
i
c

F
o
S

=

1
T
r
a
d
i
t
i
o
n
a
l

a
p
p
r
o
a
c
h

Figure 12.15 Foundation and superstructure actions under various design
approaches.

Table 12.2 Foundation displacements under various design approaches.
___________________________________________________
Design approach Traditional Seismic Yielding
FoS=1 foundation
___________________________________________________
Static factor of safety 16 8 3
Elastic sliding (mm) -10 to 12 -19 to 18 -17 to 17
Elastic rotation (mrad) -1 to 1 -3 to 3 -7 to 7
Residual sliding (mm) 0 0 2
Residual rotation (mrad) 0 0 0.5
Settlement (mm) 0 0 0.5
___________________________________________________

As discussed above, under the traditional design approach the system is
subjected to high spectral acceleration. This causes both the foundation and
the superstructure to attract more load than under the other two design
approaches. The reason why such a high static factor of safety is required to
prevent any yielding during this particular earthquake is that as the foundation
size is increased, it becomes stiffer and attracts more load. This in turn requires
the foundation to be sized even larger to prevent bearing failure from
occurring under the higher attracted load. Significant inefficiencies can result
from this process of chasing ones tail.

Under the design approach where the foundation is permitted to yield during
an earthquake, the system is comparatively much less stiff, with a longer natural
period. It is subjected to lower spectral acceleration, and as a result, the
foundation and superstructure both attract less than half the loads of those in
the traditional design approach.

Note that behaviour different to that discussed could occur for earthquakes
with different shaped response spectra.
Earthquake Resistant Design of Foundations


26

Permanent foundation displacements
Although the yielding foundation design approach may have the benefit of
lower foundation and structural loading when compared with the traditional
design approach, there is of course the consequence of accumulated permanent
foundation displacement as a result of the instances of bearing failure. Table
12.2 presents the range of elastic displacements and the residual plastic
(permanent) displacements for each of the three design approaches.

The permanent displacements under the yielding foundation design approach
are quite modest, especially when compared with the elastic displacements.
Therefore, the question must be asked: is it time to reassess the traditional
design approach for earthquake resistant shallow foundations? It is proposed
that a yielding foundation that experiences some permanent displacement may
be acceptable if the displacements are modest, and if significant benefits can be
obtained in terms of reduced foundation and superstructure actions.

12.4.2 Design criteria for yielding foundations

This section uses numerical modelling to demonstrate the type of foundation
behaviour that may be expected in relation to each of the performance aspects
discussed in the previous section. The results of the modelling are used to
discuss how criteria may be developed for each aspect. For the modelling, the
foundation from the example discussed above with a static factor of safety of
three was used. Table 12.3 presents the soil stiffness and damping values that
were calculated for the foundation using formulae presented by Gazetas
(1991). The dynamic inputs for the modelling were 20 different earthquake
records. For the purposes of illustration, and to capture a wide range of
foundation responses, the earthquake records were selected indiscriminately,
and do not represent a single type of site condition.
Foundation rotation and sliding
Figure 12.16 presents scatter plots of the maximum absolute total (elastic plus
plastic) rotation experienced by the foundation at any time during each of the
earthquakes. The top plot presents the data against the horizontal peak ground
acceleration of the earthquake, while the bottom plot presents the data against
the spectral acceleration of the earthquake at the system natural period
(approximately 1.35 seconds, Figure 12.14). Similar results were obtained for
sliding. The rotation exhibits no correlation with horizontal peak ground
acceleration, but very good correlation with the earthquake spectral
acceleration at the system natural period. This reinforces the idea that the
entire soil-foundation-structure system should be considered in design, as each
component of the system has an effect on its dynamic characteristics and
natural period. It may be possible to develop guidelines that require the system
natural period to be sufficiently different to that of the design earthquake.
Chapter 12: Soil-Foundation-Structure-Interaction

27

0
10
20
30
40
0 0.5 1 1.5 2
Earthquake spectral acceleration at Tn (g)
M
a
x
i
m
u
m

a
b
s
o
l
u
t
e

t
o
t
a
l

r
o
t
a
t
i
o
n

(
m
r
a
d
)
0
10
20
30
40
0 0.5 1 1.5 2
Horizontal Peak Ground Acceleration (g)
M
a
x
i
m
u
m

a
b
s
o
l
u
t
e

t
o
t
a
l

r
o
t
a
t
i
o
n

(
m
r
a
d
)

Figure 12.16 Residual foundation rotation versus PGA and spectral
acceleration at system natural period.

Table 12.3 Foundation stiffness and damping parameters for static factor of
safety of three.
___________________________________________________
Foundation horizontal stiffness 3.3 10
5
kN/m
Foundation rotational stiffness 1.2 10
7
kNm/rad
Foundation vertical stiffness 5.0 10
5
kN/m
Foundation horizontal damping 1.6 10
4
kN.s/m
Foundation rotational damping 3.5 10
4
kN.m.s/rad
Foundation vertical damping 3.5 10
4
kN.s/m
___________________________________________________

Figure 12.17 elaborates on Figure 12.16 by separating out the different
components of foundation rotation. The components considered are the
maximum absolute elastic, plastic, and total rotation experienced by the
foundation at any time during the earthquakes, and the residual plastic rotation
that remained at the end of the earthquakes. Similar results were obtained for
sliding, and the points noted below are equally applicable to sliding:
In this example, foundation bearing failure and associated plastic rotation
do not occur below a spectral acceleration of approximately 0.2g;
during the earthquake, maximum elastic rotation is higher than maximum
plastic rotation for spectral accelerations up to approximately 0.4g to 0.5g.
At such levels of spectral acceleration, performance during an earthquake
may be a more critical criterion than post-earthquake residual
displacements, and therefore allowing a foundation to yield may not result
in performance worse than that of a non-yielding foundation;
Earthquake Resistant Design of Foundations


28

0
5
10
15
20
25
30
35
40
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Earthquake spectral acceleration at T
n
(g)
R
o
t
a
t
i
o
n

(
m
r
a
d
)
Maximum absolute total rotation
Maximum absolute plastic rotation
Maximum absolute elastic rotation
Absolute residual rotation

Figure 12.17 Elastic and plastic foundation rotation versus spectral
acceleration at system natural period.

if bearing failure occurs, the maximum elastic rotation is virtually constant
regardless of the spectral acceleration. This is because the maximum
foundation actions (shear and moment) are limited by the bearing strength
of the soil;
conversely, if bearing failure occurs, maximum plastic rotation during an
earthquake increases roughly linearly with spectral acceleration, which is a
good indicator of foundation performance;
the maximum plastic rotation does not always equal the residual rotation, as
plastic rotation can occur in both directions during an earthquake;
therefore, the residual rotation obtained from the numerical model does not
reliably represent the maximum possible residual rotation of the foundation.
In many cases, the calculated residual rotation following the earthquake is
less than the maximum plastic rotation that occurs during the earthquake,
due to the characteristics of the particular earthquake records used;
in reality, the actual residual rotation could lie anywhere within the range of
plastic rotations that are experienced, so the calculated maximum plastic
rotation during the earthquake may be a better indicator of residual rotation
for design purposes.
Foundation settlement
Figure 12.18 presents the calculated earthquake-induced foundation settlement
versus spectral acceleration at the system natural period, for all of the 20
earthquake records used.

Chapter 12: Soil-Foundation-Structure-Interaction

29

0
50
100
150
200
0 0.5 1 1.5 2
Earthquake spectral acceleration at T
n
(g)
E
a
r
t
h
q
u
a
k
e
-
i
n
d
u
c
e
d

s
e
t
t
l
e
m
e
n
t

(
m
m
)
Elastic settlement
under static load

Figure 12.18 Earthquake-induced foundation settlement versus spectral
acceleration at system natural period.

There is a non-linear (approximately parabolic) relationship between settlement
and spectral acceleration. As a result, earthquake-induced settlements are zero
or negligible at low spectral accelerations. The static elastic foundation
settlement of approximately 50 mm is higher than earthquake-induced
settlements for spectral accelerations below approximately 0.9g, which would
usually correlate to a very strong earthquake. In fact, of the 20 earthquakes
records used, only one caused settlement greater than the static elastic
settlement.

Design criteria could compare static and earthquake-induced settlement to
assess the significance of the earthquake-induced settlement. The results
suggest that earthquake-induced settlements are only significant under very
unfavourable conditions where strong spectral accelerations coincide with the
natural period of a soil-foundation-structure system.

12.4.3 Effect of variability of soil shear strength on macro-
element modelling

In November 2009 an international workshop was held in Auckland on Soil-
Foundation-Structure-Interaction (Orense et al 2010). Some of the attendees
proposed that the beneficial effects of allowing brief instances of bearing
strength failure beneath shallow foundations during an earthquake could
outweigh the negative effects, as discussed in the previous section. During a
discussion session the issue was raised that natural soil properties may be so
variable that this performance based design approach might not be reliable.

This section examines the implications of natural variability in the undrained
shear strength and stiffness of a cohesive soil on the earthquake response of a
shallow foundation. Calculating the response of a simple structure-foundation
system to seven different earthquake records to show the implications of soil
variability for the maximum actions imposed on a shallow foundation, the
Earthquake Resistant Design of Foundations


30

natural period of the soil-foundation-structure system, and permanent
foundation displacements.

The example scenario is hypothetical, but represents a simplified structural type
that has been used elsewhere to demonstrate soil-foundation-structure-
interaction effects, for example by Paolucci et al (2011). The structure-
foundation system was set at a Class C (shallow soil) site in Wellington. With
adequate shear strength this type of site would usually be suitable for shallow
foundations. A 1000 year return period earthquake was selected as the design
earthquake, which is applicable to high value structures with a design life of
50 years.

It was assumed that the site was underlain by homogeneous cohesive soil with
a mean undrained shear strength of 100 kPa, which is typical of a Class C site.
A log-normal distribution was adopted to represent soil strength variability
following the examples given by Fenton & Griffiths (2008). The log-normal
distribution is commonly used when considering soil variability; in this case it is
required to avoid negative values of undrained shear strength.

Coefficient of variation (CoV) values of 10%, 30%, and 50% (standard
deviations of 10 kPa, 30 kPa, and 50 kPa respectively) were considered (Figure
12.11). Lumb (1974) indicates that typical CoV values for undrained shear
strength of clays are 20% to 50% and Baecher and Christian (2003) also give
values in the range of 20% to 50%. This means that the range of CoV values
chosen for the modelling herein covers the range of variability that can be
expected of real soil deposits.

Ultimate bearing pressure was assumed to equal 6s
u
. The Eurocode 8, Part 5
(BSI, 2005) undrained bearing strength surface shows that, for the bearing
strength mobilised under the static loading imposed by the structure-
foundation system considered herein, earthquake inertial effects in the soil
beneath the foundation are negligible. Thus the Eurocode 8 surface without
earthquake effects was used to define combinations of foundation actions
(moment, shear, and vertical load) that would cause bearing failure. The macro-
element model used assumed that foundation response was linear elastic when
actions are inside the bearing strength surface and perfectly plastic when
actions reached the surface (when bearing failure occurred).

The structure considered was the same bridge pier specified Table 12.1,
supported by a shallow foundation, subject to horizontal earthquake loading
acting perpendicular to the bridge. A simplifying assumption was made that
stiffness was linearly related to undrained shear strength, with G = 100 s
u
, a
soil stiffness value that includes the PGA-related reduction specified in
Table 4.1 of Eurocode 8 Part 5 (BSI, 2005). Foundation elastic stiffness and
radiation damping values were calculated using formulae presented by Gazetas
(1991).


Chapter 12: Soil-Foundation-Structure-Interaction

31

0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1.6
0.0 0.5 1.0 1.5 2.0
T
1
(s)
R
e
c
o
r
d

s
c
a
l
e

f
a
c
t
o
r
,

k
1
,

n
o
r
m
a
l
i
s
e
d

b
y

k
1

a
t

f
i
x
e
d

b
a
s
e

p
e
r
i
o
d
Arcelik
Duzce
El Centro
Hokkaido
La Union
Lucerne
Tabas
F
i
x
e
d

b
a
s
e

p
e
r
i
o
d

Figure 12.19 Probability density functions for undrained shear strength with
log-normal distribution.

0.00
0.01
0.02
0.03
0.04
0.05
50 75 100 125 150 175 200
Undrai ned shear strength, s
u
(kPa)
P
r
o
b
a
b
i
l
i
t
y

d
e
n
s
i
t
y

f
u
n
c
t
i
o
n
Coefficient of variation 10%
Coefficient of variation 30%
Coefficient of variation 50%
M
e
a
n

s
u


Figure 12.20 Record scale factors for the seven earthquake records plotted
against fundamental system period.

The seven earthquake records recommended for time history analyses for Class
C soil sites in Wellington (Oyarzo-Vera et al, 2010) were used. The earthquake
records were scaled according to the design code NZS1170.5:2004 (Standards
New Zealand, 2004), to approximate the design spectrum. This process
involved minimising, in a least mean square sense, the difference between the
earthquake record and the design spectrum over the range of 40% to 130% of
the fundamental period of the soil-foundation-structure system.

Because soil variability affects foundation stiffness, which in turn affects
fundamental system period, the record scale factor (k
1
) for each earthquake
varied with period, sometimes up to 50%. This is shown in Figure 12.19,
where scale factors calculated for each earthquake record are plotted. This
process was required to ensure that variability of earthquake records was
minimised as much as permitted by the design code used. The process was
rather time consuming. Because seven records were used, a family scale
factor (k
2
) applying to the entire suite of records was not required (because of

Earthquake Resistant Design of Foundations


32

the higher chance that at least one of the records exceeded the design spectrum
in the period range of interest).

Calculations used all seven earthquake records, and undrained shear strengths
ranging from 50 kPa to 200 kPa (50% to 200% of the mean) at increments of
1 kPa, Figure 12.20. This required a total of 1050 time history analyses, which
were reasonably quick to complete using the simple macro-element model,
Toh (2008). Foundation elastic stiffness and radiation damping (and therefore
system period), and bearing strength were calculated for each undrained shear
strength value. Earthquake record scale factors were also set on a case by case
basis, to ensure scaling appropriate to the system period.

Figure 12.21 shows the effect of soil variability on the calculated fundamental
period of the soil-foundation-structure system. The system period is
considerably longer than the fixed base period, ranging from about 1 second
for an s
u
of 50 kPa to about 1.9 seconds for an s
u
of 100 kPa. For a low CoV
of 10%, the range between which 95% of periods lie is relatively small at about
0.3 seconds. For a very high CoV of 50%, this range is considerably larger at
over 1 second.

Figure 12.22 presents the calculated relationship between maximum actions
supported by the foundation and soil strength. This relationship is dependent
only on the vertical foundation loads in relation to the ultimate bearing
strength. This is because, as explained above, the size of the bearing strength
surface was not dependent on the magnitude of earthquake loading, and the
foundation was sized to achieve a static factor of safety of 3 (without
considering seismic loading). The actions increase with increasing undrained
shear strength, because the foundation can support higher loads as soil
strength increases.

The maximum plastic displacements that occurred at any time during or after
the earthquake were considered to be the best indicator of maximum potential
permanent foundation displacements (Toh and Pender, 2010). The calculated
residual plastic displacement at the end of the earthquake is often less than this
maximum value, because plastic displacements can occur in both directions.
The absolute values of maximum plastic settlement and rotation of the
foundation are plotted in Figure 12.23 for all seven earthquake records. The
foundation sliding plot is not shown because it is very similar to the rotation
plot.

Chapter 12: Soil-Foundation-Structure-Interaction

33

0%
20%
40%
60%
80%
100%
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2
Fundamental peri od of soi l -foundati on-structure system, T
1
(s)
C
u
m
u
l
a
t
i
v
e

p
r
o
b
a
b
i
l
i
t
y
Coefficient of variation 10%
Coefficient of variation 30%
Coefficient of variation 50%
T
1
at mean s
u
of 100 kPa
F
i
x
e
d

b
a
s
e

p
e
r
i
o
d


Figure 12.21 Cumulative probability of fundamental system period for
different levels of soil variability.

0
2,000
4,000
6,000
8,000
10,000
12,000
50 75 100 125 150 175 200
Undrai ned shear strength, s
u
(kPa)
M
a
x
i
m
u
m

f
o
u
n
d
a
t
i
o
n

s
h
e
a
r

f
o
r
c
e
,

H

(
k
N
)
0
20,000
40,000
60,000
80,000
100,000
120,000
M
a
x
i
m
u
m

f
o
u
n
d
a
t
i
o
n

m
o
m
e
n
t
,

M

(
k
N
m
)
Shear
Moment
M
e
a
n

s
u


Figure 12.22 Relationship between maximum foundation actions and
undrained shear strength.

The displacements are a result of complex interaction between soil strength,
foundation stiffness, the natural period of the system, and the variable
characteristics of the earthquake records. Increasing soil strength and
increasing foundation stiffness have counteracting effects. Increasing soil
strength increases the bearing strength of the foundation, which decreases
foundation displacements. However, increasing foundation stiffness decreases
the system natural period, which (usually) leads to an increase in spectral
accelerations (i.e. moving up the right leg of the design spectrum), and an
associated increase in foundation displacements. The randomness of the
earthquake records adds further complexity to the calculated displacements,
and it is not possible to identify a fixed relationship between soil strength and
displacements. The following observations are made:
- The magnitudes of calculated displacements would probably be acceptable
in most circumstances.
- The range of calculated displacements from all seven earthquake records is
fairly similar across a wide range of undrained shear strengths. The range
at the mean s
u
value is fairly representative of the range observed at almost
all other values of s
u
. This indicates that the variability of the earthquake
Earthquake Resistant Design of Foundations


34

records has a more significant effect on displacements than soil strength
variability.
- There may be a slight decrease in displacements (and the range of
displacements) for high values of s
u
. Further, calculated maximum elastic
displacements during the earthquake (not shown) decreased slightly with
increasing soil strength, due to increasing foundation stiffness.
- Only for high variability associated with low s
u
values (in this case for s
u

values near 50 kPa) is there an observable increase in displacements. These
cases would have a very low static factor of safety, so there are likely to be
other more significant issues (e.g. static settlement).

0
0.02
0.04
0.06
0.08
0.1
50 75 100 125 150 175 200
S
e
t
t
l
e
m
e
n
t

(
m
)
Range of cal cul ated settl ements
at mean s
u
of 100 kPa
0
0.01
0.02
0.03
0.04
50 75 100 125 150 175 200
Undrai ned shear strength, s
u
(kPa)
M
a
x
.

p
l
a
s
t
i
c

r
o
t
a
t
i
o
n

a
t

a
n
y

t
i
m
e

d
u
r
i
n
g

o
r

a
f
t
e
r

e
a
r
t
h
q
u
a
k
e

(
r
a
d
)
Arcelik Duzce El Centro Hokkaido
La Union Lucerne Tabas Average
Range of cal cul ated rotati ons at
mean s
u
of 100 kPa

Figure 12.23 Calculated foundation displacements across the range of soil
strength considered, for all 7 earthquakes.

Chapter 12: Soil-Foundation-Structure-Interaction

35

Using the results presented in Figure 12.23, and the probability distributions
shown in Figure 12.20 and 21, foundation displacements were plotted against
the cumulative probability of the occurrence of the undrained shear strength
values associated with the displacements. This allowed comparison between
different values of CoV. These comparisons are shown in Figure 12.24, with
the plots indicating the average displacements corresponding to each CoV
considered. Also plotted are the minimum and maximum displacements from
all seven earthquake records. The following observations are made:
- Average displacements are relatively constant for cumulative probabilities
of s
u
greater than about 10-20%, for all CoV values considered. This
indicates that there is a high probability that permanent foundation
displacements will not be significantly affected by soil strength variability.
- The CoV of s
u
has little or no observable effect on average displacements.
The calculated average displacements are mostly very similar regardless of
the CoV value. The only clear significant effect is on settlement, for very
high CoV values.

0%
20%
40%
60%
80%
100%
0 0.01 0.02 0.03 0.04
Max. pl asti c rotati on at any ti me duri ng or after earthquake (rad)
C
u
m
u
l
a
t
i
v
e

p
r
o
b
a
b
i
l
i
t
y
0%
20%
40%
60%
80%
100%
0 0.02 0.04 0.06 0.08 0.1
Settl ement (m)
C
u
m
u
l
a
t
i
v
e

p
r
o
b
a
b
i
l
i
t
y
Coefficient of variation 50%
Coefficient of variation 30%
Coefficient of variation 10%
Average val ues
Mi ni mum
val ues
Maxi mum
val ues


Figure 12.24 Foundation displacements compared with the cumulative
probability of the undrained shear strength values associated with the
displacements.

Earthquake Resistant Design of Foundations


36

- The range between minimum and maximum displacement values for the
same CoV is much greater than the range between average values for
different CoVs, again demonstrating the variability of the earthquake
records is more significant than soil strength variability.


12.5 SHALLOW FOUNDATION SPRING-BED MODELS

This section considers simplified spring-bed models for representing soil-
foundation interaction; foundations that are relatively stiff are considered. As
explained in earlier chapters on the design of shallow foundations we need to
consider vertical, horizontal and rotational stiffness. Of particular importance
is the possibility that at some stage during earthquake loading the moments
applied to the foundation might lead to uplift of the edges of the footing with
consequent reduction in stiffness. An appealingly simple way of handling this
eventuality is to represent the soil beneath the foundation as a bed of
independent springs which can detach and reattach as the foundation rocks
back and forth. The main thrust of this section is concerned with the vertical
and rotational stiffness of shallow foundations. The stiffnesses obtained using
the spring-bed and elastic continuum models are compared. The main
conclusion is that, for a rigid rectangular foundation resting on the ground
surface, the rotational stiffness of a bed of springs is less than that of a
continuous elastic material. Furthermore, when the soil is modelled as a
nonlinear material the ratio of rotational to vertical stiffness gradually decreases
as bearing failure is approached.

In the above paragraph the terms relatively stiff and relatively flexible are used.
These terms depend on a stiffness parameter related to the bending stiffness of
the foundation and the elastic properties of the foundation. At this point we do
not need to develop a relative stiffness parameter other than to note that the
shallow foundation can usually be considered as rigid, in the sense that any
bending deformation of the foundation itself will be small in relation to the
deformation of the underlying soil.

Set-out in chapter 9, are the expressions Gazetas (1990) and his co-workers
developed giving vertical and rotational stiffness of rigid shallow foundations
of arbitrary shape and embedment. In this chapter we consider only
foundations at the ground surface. The vertical stiffness of a rectangular
foundation, obtained by combining equations 9.18 to 9.21, is given by:


0 75
b
V elastic continuum 2 2
A GL
K 0 73 1 54
1 L
.
_
. .



= +


-n



where: K
V_elastic continuum
is the vertical elastic stiffness of the foundation, G is
the shear modulus of the soil, v is the Poissons ratio, L is the length of the
foundation, and A
b
the contact area between the underside of the foundation
and the soil below.


Chapter 12: Soil-Foundation-Structure-Interaction

37

0 5 10 15 20
0
20
40
60
80
100
Footing length (m)
R
a
t
i
o
o
f
r
o
t
a
t
i
o
n
a
l
t
o
v
e
r
t
i
c
a
l
s
t
i
f
f
n
e
s
s
Elastic continuum
Discrete springs


Figure 12.25 Ratio of the rotational to vertical stiffness of square rigid shallow
footings, for a uniform elastic foundation material and the foundation
represented as a bed of discrete springs.

The rotational stiffness about an axis through the centroid of the foundation,
reorganised from equations 9.28 and 9.33, is given by:


0 15
0 75
axis
elastic continuum
3GI L
K
1 B
.
.
_ q


=

-n


where: K
u_elastic continuum
is the elastic rotational stiffness of the foundation, I
axis

is the second moment of area of the shallow foundation about the axis of
rotation, and B is the length of the axis about which the rotation occurs.
The vertical stiffness of a rectangular foundation on a bed of springs is given
by:

V springs b s
K A k
_
=

where: K
V_springs
is the vertical stiffness of the foundation on a bed of springs
and k
s
is the spring stiffness per unit area of the foundation.

The rotational stiffness of the rectangular foundation on a bed of springs is
given by:

( ) ( )
L springs 2 L springs 2
N N
2 2
2 3
springs fs s s
j 1 j 1
1 1
K A s k 2j 1 Bs k 2j 1
2 2
_ / _ /
_ q
= =
= - = -



where: k
s
is the vertical stiffness of the bed of springs per unit area of the
foundation, s is the spring spacing (assumed to be the same in the length and
breadth directions), j is a counter, N
L_springs
is the number of rows of springs
(assumed to be even) in the longitudinal direction of the footing (= L/s), and
A
fs
is the area of the foundation associated with each row of springs (=Bxs).


Earthquake Resistant Design of Foundations


38

In Figure 12.25 the ratio of the rotational stiffness to the vertical stiffness of
square foundations resting on the ground surface is plotted for both the elastic
soil and the bed of springs. These calculations were done by setting the spring
stiffness for the bed of springs so that, for given foundation dimensions, the
vertical stiffness was the same for both models. Figure 12.25 makes very clear
that the rotational stiffness of a shallow foundation on a bed of springs is
considerably less than that when the foundation is on a continuous elastic
material.

The explanation of this difference is apparent if one considers the reaction
pressure distribution beneath these foundations. For the bed of springs at
every point the reaction pressure depends only on the displacement at that
point, so for uniform vertical displacement of a rigid foundation there will be a
uniform reaction pressure. Similarly for a rotational displacement of a
foundation on the bed of springs the reaction pressure distribution will be
linear as dictated by the foundation displacements. However, for a rigid
foundation on an elastic material the reaction pressure distribution is not
uniform for a uniform settlement. The pressure tends to be very large at the
edges and corners. The reason for this is that the strains imposed on the soil
are very large at the edges of a rigid foundation. Furthermore the pressure at
any point beneath the foundation influences the pressure at every other point
beneath the foundation. The calculated pressure distribution for uniform
vertical displacement of a rigid square footing resting on an elastic layer and is
plotted in Figure 12.26a. The vertical load applied to the foundation resting on
saturated clay with an undrained shear strength of 100 kPa is such that the
ultimate limit state LRFD inequality is satisfied (in out-of-date terminology we
would say that the bearing capacity factor of safety at this vertical load is 3).
The calculation was done using the well-known solution for the vertical
displacement of the surface of an elastic half-space when a pressure loading is
applied over a rectangular area (reproduced by Poulos and Davis 1974). In a
similar manner the pressure distribution when the foundation is subject to
moment can be calculated. The distribution when the foundation is subject to
moment superimposed on the vertical loading is plotted in Figure 12.27a. The
magnitude of the moment in this case is such that one edge of the foundation
is at the point of generating negative contact pressure, that is the edge of the
foundation is about to start pulling on the soil below. Clearly this is not
possible, and the underside will start to detach from the soil.

Thus for both vertical and moment loading of rigid shallow square foundations
on elastic soil the reaction pressure distribution is far from constant or linear,
but concentrated towards the edges. It is the concentration of the reaction
towards the edge of the foundation that is the explanation for the rotational
stiffness of the foundation on the elastic soil being so much larger than the
rotational stiffness of the same sized foundation on a bed of springs (the
stiffness of which is adjusted so that the vertical stiffness of the foundations is
the same). The underlying reason is that for the spring foundation there is no
interaction between the springs so that what happens at one point has no
communication with what happens at the other points.

Chapter 12: Soil-Foundation-Structure-Interaction

39

If we examine the distributions of reaction pressure beneath the square
foundation in Figure 12.26a and 12.27a with respect to the ultimate bearing
pressure (q
u
= 5.14x1.2xs
u
~ 617 kPa) we find that there are some locations
where the elastic pressure exceeds this value. In Figure 12.26b and 12.27b the
pressure distributions are replotted with any values in excess of 617 kPa
clipped. It is apparent that only a small number of locations near the edges of
the footing are clipped because of local bearing failure.

So far we have found that for a rigid shallow foundation there is an
incompatibility between the stiffness modelling on a continuous material and
on a bed of springs. This becomes most apparent when one considers the ratio
of the rotational stiffness to the vertical stiffness a bed of elastic springs
giving a smaller rotational stiffness than a continuous elastic medium.

If one is concerned only with elastic modelling this can be remedied by adding
an additional rotational spring beneath the footing, or alternatively adding
additional width to the footing and adjusting the spring stiffnesses so that the
vertical and rotational stiffnesses of the footing are the same as that given by
equations 1 and 2. Both of these approaches have been tried, Wotherspoon et
al 2004, Pender et al 2006. However, one can ask why go to all this trouble?
Surely the correct foundation stiffness can be obtained by using separate
discrete springs for the rotational and vertical stiffness of the footing, as is
catered for in the majority of software packages for the analysis of structures.
The attraction of the bed of springs is simply that it allows modelling of the
progressive uplift of the shallow foundation under moment loading. The quest
to achieve a correct ratio of vertical to rotational stiffness is so that any
progressive uplift from the edges of the footing is modelled realistically.

0
2
4
6
8
10
0
2
4
6
8
10
200
200
400
400
600
600
800
800
1000
1000
X (m)
Y (m)
v
(kPa)
0
2
4
6
8
10
0
2
4
6
8
10
200
200
400
400
600
600
800
800
1000
1000
X (m)
Y (m)
v
(kPa)
a
b

Figure 12.26 Pressure distribution beneath a vertically loaded square rigid
shallow foundation. (a) uniform elastic soil, (b) results in (a) clipped when
the ultimate bearing pressure of the soil exceeded.
Earthquake Resistant Design of Foundations


40

0
2
4
6
8
10
0
2
4
6
8
10
0
500
1000
1500
2000
X
(
m
)
Y
(
m
)
v
(
k
P
a
)
0
2
4
6
8
10
0
2
4
6
8
10
0
500
1000
1500
2000
X
(
m
)
Y
(
m
)
v
(
k
P
a
)
a
b

Figure 12.27 Pressure distribution beneath a square rigid shallow foundation,
subject to the same vertical load as in Figure 12.26 and moment applied
subsequently. (a) uniform elastic soil, (b) results in (a) clipped when the
ultimate bearing pressure of the soil exceeded.

In Figures 12.26b and 12.27b the pressure distributions have been clipped
when the bearing pressures have been exceeded, these diagrams do not show
the next stage of the calculation in which the load represented by the clipped
part of the pressure distribution is redistributed so the same vertical load
continues to be applied to the footing.

The rotational response of a shallow foundation resting on a continuous layer,
calculated with the three-dimensional finite element software Abaqus (Simula
(2010)) is plotted in Figure 12.28. The soil is considered to be a uniform elastic
material and the foundation is resting on the surface, but the contact between
the foundation and the underlying material is not able to sustain tensile contact
stresses so when the moment is sufficiently large uplift at one side of the
foundation occurs. As can be seen from Figure 12.28 the moment-rotation
relationship for the foundation is highly nonlinear, but, as the system is elastic,
the unloading curve follows the loading curve follows back the loading curve
exactly. The plot also indicates that for small moment the rotational stiffness
obtained from Abaqus is very close to the Gazetas value given by equations
9.28 to 9.33.

Chapter 12: Soil-Foundation-Structure-Interaction

41

-2 0 2 4 6 8 10 12 14 16 18
-40
-20
0
20
40
60
80
100
120
140
Rotation (millirads)
M
o
m
e
n
t

(
k
N
m
)


Loading
Unloading


Figure 12.28 Moment-rotation curve for a rigid foundation resting on an
elastic layer for which uplift is possible.

0 2 4 6 8 10
0
20
40
60
80
100
120
140
Rotation (millirads)
M
o
m
e
n
t

(
k
N
m
)


Gezatas formula
Abaqus
OpenSEES - constant elastic
Ruaumoko - constant elastic
OpenSees - FEMA elastic


Figure 12.29 Elastic moment-rotation curves compared.

Next we need to compare the moment rotation curves obtained from Abaqus
with those for a simple spring bed. This is also done in Figure 12.28. Finally,
included in Figure 12.29 is the moment-rotation curve obtained if the spring
bed is modified in the way suggested in the FEMA 356 document where the
stiffness of the springs is increased towards the edges of the foundation. As
expected from Figure 12.25 the simple elastic spring-bed underpredicts the
rotation stiffness of the foundation, but the version using the FEMA
provisions matches the rotational stiffness quite well.
Earthquake Resistant Design of Foundations


42


We have demonstrated that when modelling the soil beneath a shallow
foundation as a bed of discrete elastic springs, with the spring stiffness set so
that the vertical stiffness of the foundation on the bed of springs is the same as
that for a footing on an elastic soil, the rotational stiffness of the foundation on
the springs is less than that of the foundation on a uniform elastic soil. The
nature of the pressure distribution beneath the foundation on an elastic soil
provides the explanation for the differing rotational stiffnesses in the case of
the elastic soil the reactions are concentrated towards the edges of the
foundation.

Three dimensional finite element calculations with Abaqus show that the
elastic rotational stiffness given by the Gazetas expressions give accurate
modelling to the elastic rotational stiffness of the foundation for small
moments but that uplift of the foundation makes the moment-rotation curve
highly nonlinear. Also the Abaqus calculations show that the suggestion given
in the FEMA 356 document where the vertical spring stiffness is concentrated
towards the edges of the foundation gives accurate modelling of the initial
moment-rotation response.


12.6 COMPARISON OF SHALLOW FOUNDATION MODELS

Shallow foundations are, of course, feasible only in non-liquefiable soils.
Liquefiable deposits will need either ground improvement to make them
suitable for shallow foundations or, more likely, mandate the use of deep
foundations. Likewise deposits of soft sedimentary clay, normally consolidated
or lightly overconsolidated, will require deep foundations. This leaves sites with
stiff cohesive soils as contenders for shallow foundations during earthquakes
(although deep foundations could be used here also). Residual soils and
overconsolidated sedimentary clays that would be described as stiff, or even
hard, soils are thus candidates for shallow foundations.

The conventional wisdom is that shallow foundation design is controlled by
static settlement. This statement is undoubtedly true when the long term
behaviour of a foundation under static load is being considered, and
particularly when one allows for the increase in bearing strength caused by the
consolidation of the soil beneath from the stresses generated by the weight of
the building and contents. Thus for larger foundations for which the static
design is controlled by long term settlement considerations the reserve of static
bearing strength under vertical load can be expected to be generous.

However, there are two important differences when it comes to earthquake
loading. First shear and moment are applied to the foundation during cyclic
loading, consequently the bearing strength is less than that under static vertical
load. Second, since we are dealing with cohesive soils, the undrained shear
strength will control the bearing strength of the foundation during the
earthquake loading rather than the drained bearing strength which controls the
long term static bearing strength.
Chapter 12: Soil-Foundation-Structure-Interaction

43

Traditionally foundation bearing strength, in relation to the applied actions, has
been expressed in terms of a bearing strength factor of safety; this lumps into
one factor uncertainties associated with both the loading and the properties of
the soil. More recent approaches have separate factors for foundation actions
and soil properties the so-called limit state design methods. These come in
two styles - Load and Resistance Factored Design (LRFD) used in New
Zealand, Canada and parts of the United States and the Partial Factor approach
of European origin. A typical LRFD approach requires that under the design
earthquake the demand on the foundation bearing strength should not exceed
a certain fraction, say 50% to 75%, of the bearing strength. In New Zealand
LRFD Design is used for proportioning shallow foundations under earthquake
loading and the mobilisation of bearing strength is limited to about 50 to 60%,
NZS1170.5 (Standards NZ (2004) and B1/VM4 (Department of Building and
Housing (2003)). In Europe where a partial factor approach is used, the
earthquake loading of a shallow foundation is restricted to mobilising about
75% of the foundation bearing strength.

Conventional SSI leads to the conclusion that for many structures the inclusion
of foundation compliance effects will reduce the earthquake design actions
imposed on the structure. The concept is that if the period of the structure is
on the falling branch of the response spectrum then the period of the
structure-foundation system will be longer than that for the fixed base
structure and so reduce the earthquake actions (although the opposite will be
true for a rising branch of the response spectrum). However, if one couples
this thinking with the requirement that the bearing strength demand on the
foundation must satisfy the LRFD requirement, then the foundation size may
be limited by bearing strength considerations. This may require for taller
structures that the plan area of the foundation needs to be larger than the plan
area of the building. As the size of the foundation increases the rotational
stiffness increases rapidly and this reduces the amount of period lengthening
induced by the SSI effects, Pender and Butterworth (2003).

The message of the above paragraph is that driven by LRFD requirements the
shallow foundation size may be such that any advantage in system performance
from soil compliance may be offset by the large foundation required. One can
ask, however, if brief instances of shallow foundation bearing strength failure
during the course of an earthquake are actually unacceptable. The consequence
will be some permanent displacement at the end of the earthquake. Perhaps the
permanent displacement could become the foundation design criterion in just
the same way as settlement is the controlling factor in shallow foundation
design for static loads.

This suggestion is hardly radical, as this idea is well established in considering
the earthquake response of earth dams and slopes (Newmark 1965) and gravity
retaining structures (Richards & Elms 1979) in terms of the residual, or
permanent, displacement at the end of the earthquake. In taking this approach
one admits that brief instances of failure of the system during the course of the
earthquake might not be important if the permanent displacements generated
during the earthquake are modest. Perhaps one needs to extend the focus on
residual displacement to the amplitudes of the cyclic displacements during the
Earthquake Resistant Design of Foundations


44

earthquake as factors, for tall structures, such as interaction with adjacent
structures or damage to service connections. Even taking this extended view
the question still remains as to whether brief instances of bearing strength
failure during the course of an earthquake is a serious matter.

The design philosophy proposed above is not new, similar ideas have been
discussed by Taylor et al (1981), Paolucci (1997), Cremer et al (2001), Gajan et
al (2005), Ugalde et al (2007), Algie et al (2008), Anastasopoulos (2009), Pender
et al (2009) and Toh and Pender (2009). All of these papers reach the
conclusion that brief instances of bearing strength failure enhance the
performance of shallow foundations in that they reduce the earthquake actions
at the cost of modest residual deformations.

12.6.1 Example tower structure on a block foundation

To demonstrate that brief instances of bearing failure during cyclic loading may
not compromise a shallow foundation, this example compares the results
obtained from dynamic tests on a shallow foundation system in the UC Davis
centrifuge with numerical calculation of the response. The foundation system
is represented using three different macro-element models.

The elastic structure rests on shallow foundations on a layer of clay
consolidated from reconstituted San Francisco Bay Mud with an undrained
shear strength of 100 kPa. The prototype dimensions of the footing are 2.67 m
in length and 0.63 m in width. The effective height of the mass during dynamic
excitation is 4.66 m. With these footing dimensions the static bearing strength
factor of safety is 2.8. The dynamic input applied in the centrifuge was a
ramped cosine wave of frequency about 1 Hertz and building to a maximum
acceleration of 0.7 g over about 8 seconds. Further details of the centrifuge test
and results are presented by Pender et al (2009) and Rosebrook and Kutter
(2001). Figure 12.32 gives the measured response history of the centrifuge
model in terms of moment - rotation, settlement - rotation and horizontal
shear force - horizontal displacement.

Below the three different elements used to models the shallow foundation
stiffness and strength are described briefly; a feature of all three models is the
coupling between shear, moment and vertical loads. The elastic stiffness and
radiation damping of the foundations was calculated using the expressions
given by Gazetas (1991).

Foundation model 1: Bearing strength surface macro-element model
This is a version of the macro element of Paolucci (1997). A more
sophisticated variant is that of Cremer et al (2001). This model is based on a
bearing strength surface which defines the combinations of vertical load,
horizontal shear, and moment that cause bearing failure beneath a shallow
foundation. Herein the surface defined in Eurocode 8 for cohesive soils is
used; it is shown in Figure 12.10. The surface shows how the amount of
moment and shear that can be applied to a shallow foundation depends on the
vertical load. This surface acts as a yield locus, state paths inside the surface are
elastic and those on the surface perfectly plastic, a non-associated plastic
Chapter 12: Soil-Foundation-Structure-Interaction

45

potential must also be specified. Elastic behaviour of the foundation is given
by three springs the stiffness and associated radiation damping or which are
calculated using the formulae of Gazetas (1991). Further details of the model
are given by Toh (2008) and Pender et al (2009) and Toh and Pender (2009).

Foundation model 2: Spring-bed model in the Ruaumoko software
The Ruaumoko software, Carr (2003), is capable of dynamic time history
nonlinear structural analysis. One of the elements provided is a nonlinear,
compression only, detachable-reattachable spring. A bed of these springs
provides the shallow foundation model which has the facility to uplift part of
the foundation during cyclic loading and to reattach it at some stage after the
direction of motion is reversed. The springs are bilinear so yielding is possible
when the contact pressure reaches a limiting value. The details are shown in
Figure 12.30. Radiation damping is included with values calculated from the
Gazetas expressions.
Foundation model 3: Spring-bed model in the OpenSees software
This spring-bed model was developed in the OpenSees software (PEER 2009)
and is very similar to the bed of springs model developed in Ruaumoko, the
difference being that the springs in this model are non-linear, rather than
bilinear as in Ruaumoko. Forty eight q-z springs were used in the vertical
direction and one t-z spring in the horizontal direction. These non-linear spring
models, QzSimple1 and TzSimple1, were developed by Boulanger (2000).
Backbone curves for these springs have a hyperbolic shape and hysteretic
damping is generated when the direction of loading is reversed. The FEMA
356 document suggests that spring-bed models should have the springs
concentrated towards the outer ends of the foundation as shown in Figure
12.31; this was how the springs were arranged for this model. The tension
capacity of the springs was set to zero so uplift of parts of the footing could be
included.

As the three sets of calculations were done with different software, before
proceeding to the nonlinear calculations a check was made on the results
obtained when the foundation was modelled with linear elastic springs one
for the vertical stiffness, one for the horizontal stiffness, and one for the
rotational stiffness. The stiffness values for these were estimated using the
relations given by Gazetas (1991). The Youngs modulus value used in
estimating the stiffness was taken as 500s
u
, that is 50 MPa. The damping values
were also calculated from Gazetas (1991). Once it was established that all three
software packages gave the same elastic output the nonlinear response was
calculated.

Figure 12.33 presents the calculated response of the centrifuge model structure
for the three macro-element models. All responses are given at prototype scale.
Figures 12.32 and 12.33 are presented on one page to aid comparison of the
responses. The three columns of figures on the page each have the same plot,
so like outputs are found in each column. In Figure 12.33 the upper row is for
the bearing strength surface based macro-element model; the middle row in the
diagram is for the Ruaumoko spring bed model; the bottom row is for the
OpenSees spring-bed model.

Earthquake Resistant Design of Foundations


46


Bi-linear detachable
springs
Rigid beam
Excitation
Rigidbeam
Bilineardetachable
springs
Excitation
Bi-linear detachable
springs
Rigid beam
Excitation
Rigidbeam
Bilineardetachable
springs
Excitation

Figure 12.30 Ruaumoko spring-bed model.


Figure 12.31 OpenSees spring-bed model with concentration of stiffness
towards the outer edges (after FEMA 356).


It is apparent in Figure 12.33 that all three models give a reasonable
representation of the observed moment-rotation and horizontal shear-
horizontal deformation behaviour. Table 12.4 summarises the residual
displacement at the completion of the centrifuge testing and calculations.

Discussion
The first model used the bearing strength surface for cohesive soils given in
Eurocode 8. As the vertical load on the foundation was constant the actions
are constrained to a vertical section through the bearing strength surface prior
to yielding. This model produces rather boxy graphs for the moment -
rotation curve and the horizontal shear horizontal displacement plots. The
reason for this is that all behaviour within the yield locus, that is the bearing
strength surface, is elastic and nonlinear behaviour occurs only when the action
Chapter 12: Soil-Foundation-Structure-Interaction

47

Table 12.4: Comparison of prototype scale residual displacements after the centrifuge dynamic excitation
Centrifuge
measurements
Ruaumoko Macro element OpenSees
Settlement (mm) 7.1 10.2 7.7 10.1
Rotation (mrad) 0.8 0.6 4.6 0.03
Horizontal displacement (mm) 0.1 3.8 3.3 5.8
path reaches the bearing strength surface. It would be possible to make the
stiffness within the bearing strength surface degrade with the number of cycles,
or based on position within the surface as is done by Cremer et al (2001).
Another effect not accounted for in this model is uplift at the edges of the
foundation. This is the reason for the difference between the shape of the
settlement rotation plot for this model and the other two models. Uplift
could be incorporated, see for example by Cremer et al (2001), but that is an
additional complication.

The second approach used the detachable, that is no-tension, spring element
provided in Ruaumoko. We set the maximum vertical stress on any spring (that
is the load carried by the spring divided by the tributary area) to 5.14s
u
; when
this pressure is reached the spring yields and from then on the vertical stiffness
is reduced. When the direction of loading reverses then the stiffness of the
spring reverts to the original value. From Figures 12.32 and 12.33 it is
apparent that this model represents what was observed in the centrifuge
reasonably well. The moment rotation curve is plotted without the damping
contribution. When this is included the moment rotation curve thickens in
the middle and the results is very similar to that for the OpenSees spring bed
model.

The final point to make with regard to the Ruaumoko model relates to the
rotational stiffness of the bed of springs. We calculated the elastic stiffness of
our footing using the Gazetas relations. One can then determine the vertical
stiffness of the bed of springs so that it is the same as that for an elastic half
space. However, if this is done, then the rotational stiffness from the bed of
springs is considerably less than that of the half space. By adding an
additional rotational spring to the centre of the footing this deficiency is
circumvated.

In the OpenSees model, the calculated moment-rotation plot is quite similar to
the measured results. The computed settlement appears to be similar to that of
the Ruaumoko model. On the other hand the horizontal sliding range of 27
mm overpredicts the experimental sliding range of 11 mm.

Herein we have concentrated on the residual displacements at the completion
of the dynamic loading. Others interested in this approach, Cremer et al (2001),
Ugalde et al (2007) and Anastasopolous (2009), have emphasized that another
benefit of allowing brief instances of bearing failure beneath shallow
foundations during earthquake loading is that the actions applied to the
foundation and structure will be less than those on a stiffer foundation. This in
turn will lead to economies in design. However, because the foundation is less
stiff using this design approach the amplitude of the structural displacements
during the earthquake may be as important as the residual deformations, Toh
and Pender (2009).

Earthquake Resistant Design of Foundations


48




























-20 -10 0 10 20
-300
-200
-100
0
100
200
300
Rotation (mrad)
-20 -10 0 10
-10
-8
-6
-4
-2
0
Rotation (mrad)
S
e
t
t
l
e
m
e
n
t
(
m
m
)
-8 -4 0 4 8
-75
-50
-25
0
25
50
75
Displacement (mm)
H
o
r
i
z
o
n
t
a
l
s
h
e
a
r
(
k
N
)

Figure 12.32 Response measured, at prototype scale, of the model tested in the UC Davis centrifuge.
-20 -10 0 10 20
-300
-200
-100
0
100
200
300
Rotation (mrad)
-20 -10 0 10 20
-10
-8
-6
-4
-2
0
Rotation (mrad)
S
e
t
t
l
e
m
e
n
t
(
m
m
)
-8 -4 0 4 8
-75
-50
-25
0
25
50
75
Displacement (mm)
H
o
r
i
z
o
n
t
a
l
s
h
e
a
r
(
k
N
)
-20 -10 0 10 20
-300
-200
-100
0
100
200
300
Rotation (mrad)
-8 -4 0 4 8
-75
-50
-25
0
25
50
75
Displacement (mm)
H
o
r
i
z
o
n
t
a
l
s
h
e
a
r
(
k
N
)
-20 -10 0 10 20
-300
-200
-100
0
100
200
300
Rotation (mrad)
-20 -10 0 10 20
-10
-5
0
5
Rotation (mrad)
S
e
t
t
l
m
e
n
t
(
m
m
)
-8 -4 0 4 8 12 16 20 24
-75
-50
-25
0
25
50
75
Displacement (mm)
H
o
r
i
z
o
n
t
a
l
s
h
e
a
r
(
k
N
)
-20 -10 0 10 20
-10
-5
0
5
Rotation (mrad)
S
e
t
t
l
e
m
e
n
t
(
m
m
)

Figure 12.33 Prototype scale computed dynamic response of the three models to the centrifuge input motion: up
row bearing strength surface macro-element, middle row Ruaumoko spring-bed, bottom row OpenSees sp
bed model.
Chapter 12: Soil-Foundation-Structure-Interaction

49

12.7 DISPLACEMENT BASED DESIGN OF A SHALLOW
FOUNDATION

In this section a shallow foundation design example is presented. A URM wall
supported by a shallow foundation is considered. The design approach is to
use the Direct Displacement Based Design method, Priestley et al (2007).
Additional references which illustrate the application of this method to shallow
foundations are Sullivan (2010), foundations for shear walls, and Paolucci et al
(2011), shallow foundations for bridge piers. The DDBD approach is
supplemented with the data from the field push-over tests on shallow
foundations presented in Sections 12.2 and 12.3; as such the method employed
in this section is a pseudo-static one. It will be seen that design of shallow
foundations for URM walls is very dependent on the amount of foundation
rotation assumed.

The wall configuration used for the example is based on the wall shown in
Figure C7-1 of Ingham (2011), which is reproduced below as Figure 12.34.
From this diagram we will consider the foundation for a 2 storey URM wall
which is 8.4 m tall and 6.9 m in length. An important assumption through the
example is that the wall and the foundation remain in contact through the
course of the earthquake excitation. The main intention of the example is to
illustrate how, allowing foundation rotation, it is possible to arrive at a
satisfactory wall foundation system. However, to achieve this the maximum
moment applied to the foundation is greater than would be applied if the usual
LRFD requirements were imposed.

9500
4100
3200
2500
230
350


Figure 12.34 Building cross-section given as Figure C7-1 in Ingham (2011).

Earthquake Resistant Design of Foundations


50



Figure 12.35 Force displacement response of a URM pier (Figure C4-2 of
Ingham (2011)).

The central concept of DDBD is the amount of lateral displacement that is
acceptable, that is structural displacement becomes the performance measure.
Figure 12.35 reproduces part of Figure C4-2 from Ingham (2011). This shows
that URM has a very limited displacement capacity; the dark line in the diagram
is the proposal of Russell (2010). In the example calculations it is assumed that
the lateral displacement of the wall with respect to the foundation is limited of
0.4% of the wall height. With a wall of 8.4 m height this amounts of only about
34 mm. Additional displacement can be obtained by lateral and rotational
displacement of the foundation. Including these components, the displacement
at the top of the wall becomes:
0 004 u = + +
wall w f w f
u . H H u
where: H
w
is the wall height, u
f
is the foundation rotation, u
f
is the horizontal
displacement of the foundation, and u
wall
is the lateral displacement at the top
of the wall. In direct DDBD it is the lateral displacement that controls the
system period, the larger the displacement the longer the system period and the
smaller the foundation actions.

The foundation moment capacity at constant vertical load is obtained using the
bearing strength surface and solving for the moment capacity for a fixed
vertical load. The elastic stiffnesses of the foundation are obtained using the
equations presented in Chapter 9.

12.7.1Application of DDBD to proportioning of shallow
foundations

The calculation steps in the example are as follows (these are as set out in the
Mathcad document for the example and explained in more detail by Paolucci et
al (2011) and Sullivan et al (2010)):

Chapter 12: Soil-Foundation-Structure-Interaction

51

1 Obtain the design elastic displacement spectrum (NZS 1170.5 (2004))
for the site in question.
2 Specify soil properties and the wall details.
3 Specify the design displacement at the top of the wall.
4 Determine the parameters for the substitute SDOF substitute structure.
5 Specify foundation dimensions. Obtain the total vertical load carried,
static reserve of bearing strength, moment capacity of the
foundation at the applied vertical load, elastic horizontal and
rotational stiffness of the foundation.
6 Evaluate the system damping and modify the design displacement
spectrum accordingly.
7 Evaluate the period of the SDOF replacement structure, the
foundation shear, and the foundation moment. Check that
foundation moment is less than the moment capacity found in step
5.
8 Evaluate the rotational stiffness of the SDOF foundation (=
foundation moment / assumed foundation rotation).
9 Plot the foundation rotation and rotational stiffness on a diagram
where the experimentally determined normalised moment-rotation
curve for the foundation is also plotted.
10 Iterate until the foundation rotation and the rotational stiffness lie on
the experimentally determined relationship. During iteration the
parameters that can be altered are: foundation rotation, foundation
damping, soil stiffness, foundation length and width. The situation
at the end of a satisfactory iterative process is shown in Figure
12.7.3.
The calculations were done in a Mathcad document in the Appendix.
The equations used for the determining the properties of the equivalent SDOF
structure are given by Priestley et al (2007) and are reproduced here:
Characteristic displacement (Priestley et al (2007) equation 3.26):
( ) ( )
2
1 1 = =
A = A A

n n
d i i i i
i i
m / m
Equivalent mass (Priestley et al (2007) equation 3.33):
( )
1 =
= A A

n
e i i d
i
m m /
Effective height (buildings) (Priestley et al (2007) equation 3.35):
( ) ( )
1 1 = =
= A A

n n
e i i i i i
i i
h m H / m
where: m
i
and A
i
refer to the mass and displacement distributions at heights H
i

in the structure.

Use of these equations requires knowledge of the mass distribution in the
structure and an assumed displaced shape. In this example it was assumed that
2/3 of the mass is located at the first floor level and 1/3 at roof level. A linear
Earthquake Resistant Design of Foundations


52

displaced shape was assumed (ie most of the lateral deformation comes from
foundation rotation).

System damping (Priestley et al (2007) equation 3.40c):

A + A
=
A + A
found found wall wall
sys
found wall

where: A
found
and A
wall
refer to the displacement of the wall and foundation
respectively, and refers to the damping in various parts of the system. The above
equation is taken from the 2003 edition of Eurocode EC8 (CEN (2003), and
changes the spectral values relative to those for a spectrum with = 0.05 (5%).

In specifying the soil properties for the example we need the undrained shear
strength, shear modulus, and damping. In the example three values of the
undrained shear strength are used (50, 100 and 150 kPa) to give reasonable
coverage of soil conditions likely to be encountered. The small strain shear
modulus of the soil, required to find the horizontal and rotational stiffness of the
foundation at very small strains, is obtained using a correlation developed for
Auckland residual clays by Meyer (1997) that G
max
is approximately 250s
u
. The
damping values for the soil are taken from experimental data presented in Section
12. 6.

12.7.2 Results

A summary of the results obtained are given in Table 12.5. A detailed listing of the
calculations is presented in Appendix 12.1.

A point discussed further below, is that the shallow foundations are quite large,
with widths 2 to 3 m. The width is a consequence of the assumed relatively small
foundation rotation (a rotation of 0.01 radians is about 0.6 degrees).

It is apparent that the largest foundations are found when the rotation is
smallest (0.0075 radians) and for these cases the rotational stiffness of
foundation is largest. For the case when s
u
= 150 and the foundation rotation
is 0.01 radians (the result illustrated in Figure 12.36) the foundation is
narrower. However, only in the case when the foundation rotation of 0.04
radians, last line of Table 12.5, is the foundation width about the size one
would might expect for the 350 mm thick wall shown in Figure 12.34. In this
case the period of the SDOF structure is 2.13 seconds, more than double that
of the other three results. Also note that for this run the foundation damping
was set at only 15%; a larger value would be easily justified with further
reduction in foundation dimensions.

Thus it seems that the push-over based pseudo-static approach results in
foundation sizes about what one would expect only for larger foundation
rotations. The question is then: would one be happy to base the design on a
foundation rotation of say 0.04 or 0.05 radians? Going so far around the
stiffness degradation curve in Figure 12.36 suggests that there is little control
on how the foundation would perform; a small change in one of the system
Chapter 12: Soil-Foundation-Structure-Interaction

53

variables could lead to a very different response. Given that the research
undertaken by Algie (2011) was to investigate what beneficial effects
foundation rotation rocking) would have on the performance of URM
structures, the above comments, at first sight, might be disappointing. The
limitation is simply a consequence of the very restricted deformation capacity
of URM and the fact that walls are rarely taller than two storeys.

Despite this limitation, the approach presented offers a relatively simple
method for initial sizing of a shallow foundation or obtaining an estimate of
what the capacity of an existing foundation might be.

Table 12.5 Comparison of shallow foundation designs for the three soil
profiles with differing undrained shear strengths
s
u
(kPa)
u
f

(rad)
A
d
(m)
B
(m)
L
(m)
FoS
static
f
(%)
T
e
(sec)
found
M
MV
_elast
K
K
u
u

50 0.0075 0.091 3.0 8.7 4.4 40 0.90 0.65 0.14
100 0.0075 0.087 2.0 7.7 6.1 35 0.88 0.74 0.13
150 0.01 0.105 1.8 6.9 7.8 30 0.93 0.87 0.10
150 0.04 0.323 0.7 6.9 3.4 15 2.13 0.72 0.03

Figure 12.36 End-point of the iteration for DDBD of shallow a foundation for
line 3 in Table 12.5 (the foundation rotation and the normalised rotational
stiffness lie on the experimentally determined curve).





1 10
6
1 10
5
1 10
4
1 10
3
0.01 0.1 1
0
0.2
0.4
0.6
0.8
1
Foundation rotation (rads)
N
o
r
m
a
l
i
s
e
d

f
o
u
n
d
a
t
i
o
n

r
o
t
a
t
i
o
n
a
l

s
t
i
f
f
n
e
s
s

Earthquake Resistant Design of Foundations


54

12.8 INTEGRATED DESIGN OF STRUCTURE-FOUNDATION SYSTEMS

Given the very powerful computer resources that are now available for civil
engineering and infrastructure design, a pressing need is to improve interaction
between the structural and geotechnical communities. An obvious priority is
for the two groups to work together in a more integrated fashion. The author
has been promoting this view for some years; the suggestion is universally
greeted with assent. Recently, similar observations have been made with regard
to the design of very tall buildings, Poulos (2009) and Baker (2010). In
addition the FEMA 356 (Federal Emergency Management Agency (2000))
document also emphasizes that foundation and structural design need to be
considered together (Section C.4.1).

The most direct way in which this can be achieved is by the two communities
developing integrated numerical models of complete structure-foundation
systems. Too often in the past the practice has been for the foundation and
superstructure to be considered almost in isolation. Lapsing into
anthropomorphism, one can say, that, from the perspective of an incoming
earthquake, the structure and the foundation system supporting it appear as a
single entity.

If this is accepted then the design approach needs to be based on a single
integrated model of the building-foundation system. Nowadays exceedingly
capable software is used for analysis and design of structures. The full potential
of this software will not be realised until a complete model of the structure-
foundation system is used. This point of view is certainly not based on the
assumption that the future of engineering design lies in evermore sophisticated
software, in a manner that reduces human input and minimises opportunities
for engineering judgement. Rather it is intended that the exercise of
engineering design judgement will be enhanced, so enabling the designer to
obtain a more realistic understanding of the how the design will perform. This
requires little new in the way of software facilities, what is needed is simply that
the human side of the process is organised to realise the best output from the
combination of numerical modeling and geotechnical evaluation of the
materials present at a given site.

This integrated approach will be invaluable when applied to the design of new
buildings and infrastructure. However, it may be even more valuable when
applied to the assessment of existing foundations for facilities that are under
consideration for retrofit. Here careful evaluation of the manner in which a
foundation and the structure supported interact may lead to a better
understanding of the actual capacity of the system and even, in some cases, the
conclusion that retrofit is unnecessary. The reason that this could happen is
that it is likely that such assessments will be more sophisticated than the
original design and less dependent on a cascade of conservative decisions
which has been a significant feature of foundation design in the past.




Chapter 12: Soil-Foundation-Structure-Interaction

55



12.9 SUMMARY











































Earthquake Resistant Design of Foundations


56




REFERENCES

Algie, T. B. (2011) Nonlinear Rotational Behaviour of Shallow Foundations
on Cohesive Soil, Ph D thesis, University of Auckland.
Algie, T. B., Pender, M. J. and Orense, R. P. (2010) Large scale field tests of
rocking foundations on an Auckland residual soil. In Soil Foundation
Structure Interaction (R Orense, N Chouw, and M Pender (eds)), CRC
Press / Balkema, The Netherlands, pp. 57- 65.
Algie, T.B., Deng, L., Erduran, E., Kutter, B.L. & Kunnath, S. (2008).
Centrifuge Modeling of Innovative Foundation Systems to Optimize
Seismic Behavior of Bridge Structures. 14th World Conference on
Earthquake Engineering (14 WCEE).
Anastasopoulos, I (2009) Beyond conventional capacity design: towards a new
design philosophy. Proceedings of the International Workshop on Soil-
Foundation-Structure-Interaction. University of Auckland, November 26-
27, 211-218.
Baecher G.B. & Christian J.T. 2003. Reliability and Statistics in Geotechnical
Engineering. Chichester: John Wiley & Sons Ltd.
Baker, C (2010) Uncertain geotechnical truth and the cost of effective high
rise foundation design, in The Art of Foundation Engineering Practice.
ASCE.
Boulanger R.W. 2000. The TzSimple1 and QzSimple1 Material Models,
Documentation for the OpenSees Platform.
http://opensees.berkeley.edu/.
Boulanger, R.W., Curras, C.J., Kutter, B.L., Wilson, D.W. & Abghari, A.
(1999). Seismic soil-pile-structure interaction experiments and analyses.
Journal of Geotechnical and Geoenvironmental Engineering, 125(9), 750-
759.
Briaud, J-L. and Lepert, P. (1990). WAK test to find spread footing stiffness.
Journal of Geotechnical Engineering, Vol. 116, No. 3, pp. 415-431.
BSI British Standards 2005. BS EN 1998-5 Eurocode 8: Design of structures
for earthquake resistance Part 5: Foundations, retaining structures and
geotechnical aspects.
Carr, A. J. 3D RUAUMOKO: inelastic three-dimensional dynamic analysis
program, University of Canterbury - Department of Civil Engineering,
Christchurch, NZ, 2004.
Chatzigogos, C. T., Pecker, A. & Salenon, J. 2007. A macro-element for
dynamic soil-structure interaction analyses of shallow foundations. 4th Int.
Conf. on Earthquake Geotechnical Engineering, Thessaloniki, Greece,
June 25-28.
Comit Europen de Normalisation (2003) Eurocode 8, part 1: Design of
Structures for Earthquake Resistance Part 1: General Rules, Seismic
Actions and Rules for Buildings, prEN 1998-1, CEN, Brussels, Belgium.
Draft 6, January.
Comit Europen de Normalisation Eurocode 7, part 1: Geotechnical Design,
General Rules. Draft 6, January 2003.
Chapter 12: Soil-Foundation-Structure-Interaction

57

Cremer, C. and Pecker, A. and Davenne, L. Cyclic macro-element for soil
structure interaction: material and geometrical nonlinearities. International
Journal of Analytical and Numerical Methods in Geomechanics, 25(13):
1257-1284, 2001.
Deng. L., Kutter, B. L., Kunnath, S., and Algie, T. (2010) Performance of
bridge systems with nonlinear soil-footing-structure interactions. In Soil
Foundation Structure Interaction (R Orense, N Chouw, and M Pender
(eds)), CRC Press / Balkema, The Netherlands, pp. 49-56.
Federal Emergency Management Agency 2000. FEMA 356: Prestandard and
Commentary for the Seismic Rehabilitation of Buildings. Washington, D.C.
Fenton G.A. & Griffiths D.V. 2008. Risk Assessment in Geotechnical
Engineering. New Jersey: Wiley.
Gajan, S. & Kutter, B. L. (2008). Capacity, settlement, and energy dissipation
of shallow footings subjected to rocking. Journal of Geotechnical and
Geoenvironmental Engineering, 134(8): 1129-1141.
Gajan, S. & Kutter, B. L. 2007. A contact interface model for nonlinear cyclic
moment-rotation behaviour of shallow foundations. 4th Int. Conf. on
Earthquake Geotechnical Engineering, Thessaloniki, Greece, June 25-28.
Gajan, S., Kutter, B. L., Phalan, J. D., Hutchinson, T. C. & Martin, G. R.
Centifuge modeling of load deformation behaviour of rocking shallow
foundations. Soil Dynamics and Earthquake Engineering, 25: 773-783,
2005b.
Gazetas G. 1991. Foundation vibrations, Chapter 11 in Foundation
Engineering Handbook, edited by Fang H.Y. 2nd edition. New York: Van
Nostrand Reinhold.
Gazetas, G., Anastasopoulos, I., and Apostolou, M. (2007). Shallow and deep
foundations under fault rupture or strong seismic shaking. Chapter 9 in:
Earthquake Geotechnical Engineering (K. D. Pitalakis (ed.)), Springer,
The Netherlands, pp. 185-215.
Ingham, J. M. (2011) Assessment and Improvement of Unreinforced Masonry
Buildings for Earthquake Resistance. Version 02/2011. Edited by J M
Ingham.
Kramer, S. L., Arduino, P. and Shin, H. (2009) Development of performance
criteria for foundations and earth structures. Proc. of the conference on
Performance-Based Design in Earthquake Geotechnical Engineering IS-
tokyo 2009. Edited by Kokusho, Tsukmoto and Yoshimine. pp. 107-120.
Lumb P. 1974. Application of statistics in soil mechanics, Chapter 3 in Soil
Mechanics New Horizons, edited by Lee I.K. London: Newnes-
Butterworths.
Meyer, V. M (1997) Stress-strain-strength properties of an Auckland residual
soil, PhD thesis, University of Auckland.
MSaDon, N. M., Pender, M. J., Orense, R. P. and Abdul Karim, A. R. (2010)
Full-scale pile head lateral vibration tests. In Soil Foundation Structure
Interaction (R Orense, N Chouw, and M Pender (eds)), CRC Press /
Balkema, The Netherlands, pp. 33 39.
New Zealand Building Code (2008) Clause B1 - Structure: Verification
Method B1/VM4 Foundations. Department of Building and Housing,
New Zealand. Amendment 8, December. www.dbh.govt.nz
Newmark, N. (1965) "Effects of earthquakes on dams and embankments"
Geotechnique, Vol. 15 No. 2, pp 139-160.
Earthquake Resistant Design of Foundations


58

Orense R.P., Chouw N. & Pender M.J. (eds) 2010. Soil-Foundation-Structure
Interaction. London: CRC Press/Balkema, Taylor & Francis Group.
Oyarzo-Vera C., McVerry G. & Ingham J. 2010. Seismic zonation and default
suite of ground motion records for time-history analysis in the North
Island of New Zealand. Personal correspondence (submitted for
publication).
Paolucci R., Figini R., di Prisco C., Petrini L., & Vecchiottii M. 2011.
Accounting for non-linear dynamic soil-structure interaction in the
displacement-based seismic design. 5th International Conference on
Earthquake Geotechnical Engineering, Santiago 10-13 January 2011.
Paolucci, R. (1997) Simplified evaluation of earthquake-induced permanent
displacements of shallow foundations. Journal of Earthquake
Engineering. 1(3), 563-579, 1997.
Paolucci, R. 1997. Simplified evaluation of earthquake-induced permanent
displacements of shallow foundations. Journal of Earthquake Engineering
1(3): 563-579.
Pecker, A., Chatzigogos, C. T., and Salencon, J. (2010) A dynamic macro-
element for performance based design of foundations. Chapter 10 in:
Advances in Peformance-Based earthquake Engineering, Springer, pp. 103-
112.
PEER. (2009). OpenSees - Open System for Earthquake Engineering
Simulation: http://opensees.berkeley.edu/.
Pender, M J & Butterworth, J W Soil-structure interaction and the Australia
New Zealand loading standard Pacific Earthquake Conference,
Proceedings CDROM ISBN 0-473-09372-3, paper 097, 10 pages.
Pender, M. J., Toh, J. C. W., Wotherspoon, L. M. & Algie, T. B. 2009.
Earthquake induced permanent displacements of shallow foundations
performance based design. Int. Conf. of Performance-Based Design in
Earthquake Geotechnical Engineering, Tsukuba, Japan, June 15-17.
Pender, M. J., Wotherspoon, L. M. & Toh, J. C. W. 2008. Foundation stiffness
estimates and earthquake resistant structural design. The 14th World
Conference on Earthquake Engineering, Beijing, China, October 12-17.
Pender, M.J., Toh, J.C.W., Wotherspoon, L.M., Algie, T.B. & Davies, M.C.R.
(2009). Earthquake induced permanent displacements of shallow
foundations performance based design. IS Tokyo 2009, Tokyo, Japan.
Poulos, H. G. (2009) Significance of interaction non-linearity in piled
foundation raft design Proceedings of the International Workshop on
Soil-Foundation-Structure-Interaction. University of Auckland, November
26-27, 193-199.
Priestley, M. J. N., Calvi, G. M. and Kowalsky, M. J. (2007) Displacement-
Based Seismic Design of Structures. IUSS Press, Pavia.
Raychowdhury, P. & Hutchinson, T.C. (2009). Performance evaluation of a
nonlinear Winkler-based shallow foundation model using centrifuge test
results. Earthquake Engineering and Structural Dynamics, 38(5), 679-698.
Richards, R. and Elms, D. (1979) "Seismic behavior of gravity retaining walls"
Journal of the Geotechnical Engineering Division, ASCE, Vol. 105, No.
GT4, pp 449-464.
Rosebrook, K.R., Kutter, B.L. 2001. Soil-foundation-structure interaction:
shallow foundations. Centrifuge Data Report for test series KRR03,
University of California, Davis, CA, Report No. UCD/CGMDR-01/11.
Chapter 12: Soil-Foundation-Structure-Interaction

59

Russell, A. (2010) Characterisation and Seismic Assessment of Unreinforced
Masonry Buildings, PhD thesis, University of Auckland, Auckland.


M.Sadon, N (2011) Full scale static and dynamic lateral loading of a single pile.
PhD thesis, University of Auckland
Simulia 2010. Abaqus 6.8-EF2. Dessault Systemes.
Standards New Zealand (2004) NZS 1170.5:2004 Structural design actions
Part 5: Earthquake Design Actions New Zealand. Wellington.
Standards New Zealand, Structural Design Actions AS/NZS 1170.1:2002.
Standards New Zealand, The Design of Concrete Structures, NZS 3101:2006.
Stokoe, K. H., Sung-Ho, Joh., and Woods, R. D. (2004). Some contributions to
in situ geophysical measurements to solving geotechnical engineering
problems. Proc. International Conference on Site Characterisation (ICS-2),
Porto Portugal, September.
Stokoe, K. H., Wright, S. G., Bay, J. A. & Roesset, J. M. (1994).
Characterization of geotechical sites by SASW method. Geophysical
characterization of sites, pp. 15-25.
Sullivan, T. J., Salawdeh, S., Pecker, A., Corigliano, M. and Calvi, G. M. (2010).
Soil-foundation-structure interaction considerations for performance-based
design of RC wall structures on shallow foundations. In Soil Foundation
Structure Interaction (R Orense, N Chouw, and M Pender (eds)), CRC
Press / Balkema, The Netherlands, pp. 193-200.
Taylor, P.W., Bartlett, P.E. & Wiessing, P.R. (1981). Foundation Rocking
Under Earthquake Loading. Proceedings of the 10th International
Conference on Soil Mechanics and Foundation Engineering, 3, 313-322.
Toh J.C.W. & Pender M.J. 2010. Design approaches and criteria for
earthquake-resistant shallow foundation systems. In Soil-Foundation-
Structure Interaction, edited by Orense R.P., Chouw N. & Pender M.J.
London: CRC Press/Balkema, Taylor & Francis Group.
Toh J.C.W. 2010. Performance Based Aseismic Design of Shallow
Foundations. Germany: LAP Lambert Academic Publishing.
Toh, J. C. W. & Pender, M. J. (2009) Design approaches and criteria for
earthquake-resistant shallow foundation systems. Proceedings of the
International Workshop on Soil-Foundation-Structure-Interaction.
University of Auckland, November 26-27, 179-186.
Toh, J. C. W. (2008) Performance based aseismic design of shallow
foundations, ME thesis, University of Auckland.
Toh, J. C. W. 2008. Performance based aseismic design of shallow foundations.
ME thesis, University of Auckland.
Toh, J. C. W. and Pender, M. J. (2010). Design approaches and criteria for
earthquake-resistant shallow foundation systems. In Soil Foundation
Structure Interaction (R Orense, N Chouw, and M Pender (eds)), CRC
Press / Balkema, The Netherlands, pp. 173 180.
Toh, J. C. W., Pender, M. J. and McCully, R (2011). Implications of soil
variability on performance based shallow foundation design. Proc. 9th
Pacific Conference on Earthquake Engineering, Auckland, April 14-16.
Ugalde, J.A., Kutter, B.L., Jeremic, B. & Gajan, S. (2007). Centrifuge Modeling
of Rocking Behavior of Structures on Shallow Foundations. Fourth
Earthquake Resistant Design of Foundations


60

International Conference on Earthquake Geotechnical Engineering,
Thessalonaki, June 25-28..
Wotherspoon, L. M. and M. J. Pender (2010) Effect of uplift modeling on the
seismic response of shallow foundations. Proceedings 5th International
Conference on Earthquake Engineering and Soil Dynamics and
Symposium in honor of Professor I M Idriss. San Diego, May.
Wotherspoon, L. M., S. Sritharan and M. J. Pender (2010)a An Investigation
on the Impact of Seasonally Frozen Soil on Seismic Response of Bridge
Columns accepted for publication ASCE Journal of Bridge Engineering.
Wotherspoon, L. M., S. Sritharan and M. J. Pender (2010)b Modelling the
response of cyclically loaded bridge columns embedded in warm and
seasonally frozen soils accepted for publication Journal of Engineering
Structures.





































Appendix 12.1 Shallow foundation DDBD example
File: Shallow foundation DDBD example February 2011 23/01/2011- 08/02/2011
Design example for a two storey URM wall structure. Based on the 2 storey URM structure
illustrated in Fig. C7-1 of Ingham.
ORIGIN 0 :=
Counter for vector elements etc
g 9.81 :=
Gravitational acceleration (m/sec
2
)
NZS1170.5 spectral shape factor for site class D (Figure 3.1) General (not modal analysis etc)
DeepA TT ( )
EQ
i
3.00 TT
i
0.56 s if
2.4
0.75
TT
i
|

\
|
|
.
0.75

(
(
(

0.56 TT
i
< 1.5 s if
2.14
TT
i
1.5 TT
i
< 3 s if
6.42
TT
i
( )
2
TT
i
3 > if

i 0 length TT ( ) 1 .. e for
EQ
:=
i 0 100 .. := ST
i
i
4
100
:=
Vector of spectral periods (sec)
Z 0.4 :=
Wellington, or thereabouts
R 1 :=
500 year return period
N 1 :=
Site more than 20 km from a fault
(Omit the performance factor and the ductility as DDBD starts with the elastic response spectrum.
Section 2.2.3 of Priestley et al).
Spec_acc Z R N DeepA ST ( ) :=
Spectral acceleration (elastic) (g)
Displacement spectrum
Displacement spectrum derived from the above acceleration spectrum.
Spec_dis Spec_acc ST
2

( )

g
4
2

( )
:=
0 1 2 3 4
0
0.2
0.4
0.6
0.8
Spec_dis
ST
Soil properties:
s
u
150 := Undrained shear strength (kPa)
1.8 := Soil density (tonnes/m
3
)
0.5 := Clay Poisson's ratio
G
max
250 s
u
:= G
max
3.75 10
4
= Small strain shear modulus (Vaughan Meyer relationship)
V
s
G
max

:= V
s
144.338 = Shear wave velocity implied by the G
max
value.
Wall and foundation details:
W 1000 := Vertical load carried by wall foundation (kN)
M
W
g
:= M 101.937 = Effective wall mass (tonnes)
D
f
1 := Foundation depth (m)
h 8.4 := Wall height (m) (Figure C7-1 Ingham
H
w
h D
f
+ := H
w
9.4 = Wall height above foundation level (m)
Displacement performance requirements for the URM wall (URM details from the Ingham document):

f
0.01 := Assumed foundation rotation (radians)

w
0.40.01 H
w
:=
w
0.038 = Displacement from wall deformation (m)
(Figure C4-2, Ingham)

w
H
w

f
+ 0.011 + := 0.143 = Total displacement at top of wall at 0.4% drift.
(The last value in this equation is inserted once
we have been nearly to the bottom and calculated
the foundation horizontal displacement.)
Substitute structure (Chapter 3, Priestley et al (2007)):
From Fig. C7-1 let's say we have a mass of 2m at a height of 4.1 m and another mass of m at a
height of H
w
.
Assume that the above displacement, , is distributed linearly over the wall height (a slight
simplification as it neglects the small horizontal displacement at foundation level).
m
M
3
:=

d
2 m
4.1
H
w

|

\
|
|
.
2
m
2
+
2 m
4.1
H
w

|

\
|
|
.
m +
:=
d
0.105 =
Priestley et al equation 3.26
m
e
2 m
4.1
H
w

|

\
|
|
.
m +

d
:= m
e
86.287 =
Priestley et al equation 3.33
h
e
2 m
4.1
H
w

|

\
|
|
.
4.1 m H
w
+

(
(

2 m
4.1
H
w

|

\
|
|
.
m +

(
(

:= h
e
6.931 =
Priestley et al equation 3.35
Foundation dimensions and total weight carried:
L h
e
0.0 + := L 6.931 =
Foundation length (m) (Need foundation to be
longer than wall height for bearing strength)
B 1.8 :=
Foundation width (m) (iterate until appropriate Vbar
and other values)
W W B L D
f
24 + := W 1.299 10
3
=
Total vertical load on the foundation (kN) (Need
foundation weight for wall stability.)
atan
L
2
h
e
|

\
|
|
|
.
:= 0.464 =
Wall topling angle (radians) (for use of Tom's
snap-back data)
Now, need vertical load that will give bearing strength with zero shear and moment
Use Equation 4.5 as in Example 4.1:
V
uo
s
u
5.14 1 0.2
B
L
+
|

\
|
|
.
1 0.4
D
f
0.0
B
+
|

\
|
|
.
D
f
+

(
(

L B := V
uo
1.013 10
4
=
Vbar
W
V
uo
:= Vbar 0.128 = Vbar
1
7.792 =
Static factor of safety
Now need the ultimate moment capacity of the foundation, MV, with the fixed vertical load, W, and
the foundation shear is MV/h
e
. (cf Example 4.1)
The inclination factor,
ci
equation I5 in the Appendix to Chapter 4, can cause problems in two ways. First, the term
inside the square root can become negative which indicates that the foundation does not have adequate sliding
resistance. Second, the term inside the square root bracket may be greater than 1.0, whoch means that
ci
will be
greater than 1.0
Both of these are checked below. For the second
ci
is set to 1.0, deciding if this is the case is the reason for the two
values MV1 and MV2 are calculated below, an if statement decides which to use. For the first the horizontal shear
applied to the foundation is reduced by subtracting the available passive resistance in front of the foundation. The
need for this is shown in second of the checks. In fact, it is this second condition that gives rise to the first problem
which arises when 2 s
u
B D
f
is greater than M
u
/h
e
.
M
u
500 := Trial value for the root finder
MV1 root W s
u
5.14 1 0.2
L 2
M
u
W

B
+
|

\
|
|
|
.

1
2
1 1
M
u
h
e
2 s
u
B D
f
1.0
s
u
B L 2
M
u
W

|

\
|
|
.

(
(
(
(
(

(
(
(
(
(

D
f
+

(
(
(
(
(

B L 2
M
u
W

|

\
|
|
.

(
(
(
(
(

M
u
,

(
(
(
(
(

:=
MV1 3.932 10
3
=
MV2 root W s
u
5.14 1 0.2
L 2
M
u
W

B
+
|

\
|
|
|
.
D
f
+

(
(
(

B L 2
M
u
W

|

\
|
|
.

(
(
(

M
u
,

(
(
(

:=
MV2 3.948 10
3
=
MV if 1
MV1
h
e
2 s
u
B D
f
1.0
s
u
B L 2
MV1
W

|

\
|
|
.

1 > MV2 , MV1 ,

(
(
(
(
(

:= Decision whether to use MV1 or MV2


MV 3.932 10
3
= Ultimate moment capacity of the foundation at the
fixed vertical load W and shear = MV/h
e
.
Let's check the various components of the above equation to make sure we have not jumped to too
large a value for the solution:
L 2
MV
W
0.879 = Effective length of footing at moment capacity (m), OK as positive.
Check that the dimension B is larger than L', so
that the dimension B is on the bottom line of
cs
Check if L 2
MV
W
B < "OK" , "NO" ,
|

\
|
|
.
:= Check "OK" =
Thus the foundation shear cannot be taken totally by
the underside of the foundation, need to rely on some
passive pressure generated by embedment, so do not
used depth factors in the bearing strength assessment.
1
MV
h
e
2 s
u
B D
f
0.0
s
u
B L 2
MV
W

|

\
|
|
.

1.392 =
Evaluate small strain horizontal and rotational stiffnesses of the foundation:
1.0 :=
Possible modifier to the small strain soil stiffness
G G
max
:=
K
h
G L
2
2 2.5
L B
L
2
|

\
|
|
.
0.85
+

(
(
(

:=
Horizontal stiffness (kN/m) (Equations 3.10 and 3.11)
(Have not used depth terms for these as have not
used depth factors when assessing bearing strength).
Foundation rocking stiffness (kNm/rad) (Equations 3.17
and 3.18)
K

G
B L
3

12
|

\
|
|
.
0.75

1
2.4 0.5
B
L
+
|

\
|
|
.
:=
System damping:
URM wall damping (page ?? Jason)

wall
0.15 :=
Assumed foundation damping (typical value from
Tom Algies's snap-back tests, cf Fig. 6.10)

found
0.30 :=

sys

found
h
e

f

wall

w
+
h
e

f

w
+
:=
sys
0.247 =
Priestley et al 2007 equation 3.40c (Could add to the top and
bottom lines the foundation displacement the small contribution
that comes from lateral deformation.)
Adjust displacement spectrum for this damping value:
Spec_acc
sys
Spec_acc
0.10
0.05
sys
+
|

\
|
|
.
0.5
:=
Adjusted using the EC8 equation (Priestley et al
equation 2.9)
Spec_dis
sys
Spec_acc
sys
ST
2

|
\
|
.

g
4
2

( )
:=
Blue is the design displacement spectrum adjusted for the system damping.
0 1 2 3 4
0
0.2
0.4
0.6
0.8
Spec_dis
Spec_dis
sys
ST
Need to find the period associated with the system displacement
d
Locate the points either side of
d

J Count1 0
j 1
j j 1 +
Spec_dis
sys
j

d
< while
j
:=
J 24 =
Linear interpolation between positions J and J-1 to get period of equivalent SDOF system:
T
e
Spec_dis
sys
J

d

Spec_dis
sys
J
Spec_dis
sys
J 1

ST
J
ST
J 1

( )
ST
J 1
+ :=
T
e
0.926 = SDOF period (secs)
SDOF stiffness (kN/m)
K
eff
4
2

M
T
e
2
:= K
eff
4.693 10
3
=
V
b
K
eff

d
:= V
b
493.436 = Base shear (kN)
M
found
V
b
h
e
:= M
found
3.42 10
3
= Foundation moment (kNm)
MV 3.932 10
3
=
M
found
MV
0.87 =
u
V
b
K
h
1.019 10
3
= := Foundation lateral elastic
displacement from shear.
Normalised foundation variables:
Vbar 0.128 = Mbar
M
found
V
uo
L
:=
(Note the use of L rather than B, as rotation about the
breadth axis)
Check that the dimension B is larger than L', so
that the dimension B is on the bottom line of
cs
Check if L 2
M
found
W
B < "OK" , "NO" ,
|

\
|
|
.
:= Check "OK" =
Hbar
V
b
V
uo
:= Lbar
L 2
M
found
W

L
:=
Vbar 0.128 = Hbar 0.049 = Mbar 0.049 = Lbar 0.241 =
K
found
M
found

f
:= K
found
3.42 10
5
=
equivalent foundation rotational
stiffness (kNm/rad)
Foundation elastic rocking stiffness (kNm/rad)
K
.elastic
G
max
B L
3

12
|

\
|
|
.
0.75

1
2.4 0.5
B
L
+
|

\
|
|
.
:=
K
.elastic
3.564 10
6
=
K
found
K
.elastic
0.096 = 1 =
T
e
0.926 =
M
found
MV
0.87 =
Import Tom Algie's shallow foundation normalised moment-rotation curves obtained from snap-back tests:
Normal
...\Out1.xls
:=
Column 0: Normalised rotation
Column 1: Normalised fitted moment
Column 2: Normalised secant modulus

fn

f

:=
fn
0.022 =
Normalised foundation rotation.
We need a couple of ploting parameters to help indicate when we have iterated to the correct
solution:
Frot
0.6
0

f
|

\
|
|
|
.
:= Kdeg
10
6
1
K
found
K
.elastic
K
found
K
.elastic
|

\
|
|
|
|
|
|
.
:=
Above we calculated the elastic lateral displacement of the foundation under the shear developed.
However, the foundation rotational stiffness is degraded as shown by Tom's data. We need to do a
similar degradtion of the horizontal stiffness and so get the horizontal displacement corresponding to
the foundation rotation.
u
K
found
K
.elastic
0.011 =
(Include this displacement above when setting the design
displacement limit, . This requires some looping back /
iteration.)
1 10
6
1 10
5
1 10
4
1 10
3
0.01 0.1 1
0
0.2
0.4
0.6
0.8
1
Foundation rotation (rads)
N
o
r
m
a
l
i
s
e
d

f
o
u
n
d
a
t
i
o
n

r
o
t
a
t
i
o
n
a
l

s
t
i
f
f
n
e
s
s

1 =

f
0.01 =
s
u
150 =
B 1.8 =
L 6.931 =
W 1.299 10
3
=
Vbar
1
7.792 =

d
0.105 =

w
0.038 =

wall
0.15 =

found
0.3 =
T
e
0.926 =
M
found
MV
0.87 =
K
found
K
.elastic
0.096 = Vbar 0.128 = Hbar 0.049 = Mbar 0.049 =
Note that the fixed base period of the URM wall is not needed.
Increasing the foundation damping lowers the green line slightly.
Results

Case1:s
u
=50kPa(G=250s
u
)


(Include this displacement in setting the displacement limit.)











u
K
found
K
.elastic
0.015
1 10
6
1 10
5
1 10
4
1 10
3
0.01 0.1 1
0
0.2
0.4
0.6
0.8
1
Foundation rotation (rads)
N
o
r
m
a
l
i
s
e
d

f
o
u
n
d
a
t
i
o
n

r
o
t
a
t
i
o
n
a
l

s
t
i
f
f
n
e
s
s

1

f
7.5 10
3

s
u
50
B 3
L 8.731
W 1.629 10
3

Vbar
1
4.427

d
0.091

w
0.038

wall
0.15

found
0.4
T
e
0.9
M
found
MV
0.65
K
found
K
.elastic
0.14 Vbar 0.226 Hbar 0.063 Mbar 0.05
Case2:s
u
=100kPa(G=250s
u
)

(Include this displacement in setting the displacement limit)













u
K
found
K
.elastic
9.888 10
3

1 10
6
1 10
5
1 10
4
1 10
3
0.01 0.1 1
0
0.2
0.4
0.6
0.8
1
Foundation rotation (rads)
N
o
r
m
a
l
i
s
e
d

f
o
u
n
d
a
t
i
o
n

r
o
t
a
t
i
o
n
a
l

s
t
i
f
f
n
e
s
s

1

f
7.5 10
3

s
u
100
B 2
L 7.731
W 1.371 10
3

Vbar
1
6.103

d
0.087

w
0.038

wall
0.15

found
0.35
T
e
0.879
M
found
MV
0.737
K
found
K
.elastic
0.127 Vbar 0.164 Hbar 0.054 Mbar 0.049
Case3:s
u
=150kPa(G=250s
u
)

(Include this displacement in setting the displacement limit)

















u
K
found
K
.elastic
0.011
1 10
6
1 10
5
1 10
4
1 10
3
0.01 0.1 1
0
0.2
0.4
0.6
0.8
1
Foundation rotation (rads)
N
o
r
m
a
l
i
s
e
d

f
o
u
n
d
a
t
i
o
n

r
o
t
a
t
i
o
n
a
l

s
t
i
f
f
n
e
s
s

1

f
0.01
s
u
150
B 1.8
L 6.931
W 1.299 10
3

Vbar
1
7.792

d
0.105

w
0.038

wall
0.15

found
0.3
T
e
0.926
M
found
MV
0.87
K
found
K
.elastic
0.096 Vbar 0.128 Hbar 0.049 Mbar 0.049
Comparison:
s
u
(kPa)
f
(rad)
d
(m) B(m) L(m) FoS
static

f
(%) T
e
(sec)

50 0.0075 0.091 3.0 8.7 4.4 40 0.90 0.65 0.14


100 0.0075 0.087 2.0 7.7 6.1 35 0.88 0.74 0.13
150 0.01 0.105 1.8 6.9 7.8 30 0.93 0.87 0.10

Вам также может понравиться